Individual Dynamical Masses of Ultracool Dwarfs*

and

Published 2017 August 2 © 2017. The American Astronomical Society. All rights reserved.
, , Citation Trent J. Dupuy and Michael C. Liu 2017 ApJS 231 15 DOI 10.3847/1538-4365/aa5e4c

Download Article PDF
DownloadArticle ePub

You need an eReader or compatible software to experience the benefits of the ePub3 file format.

0067-0049/231/2/15

Abstract

We present the full results of our decade-long astrometric monitoring programs targeting 31 ultracool binaries with component spectral types M7–T5. Joint analysis of resolved imaging from Keck Observatory and Hubble Space Telescope and unresolved astrometry from CFHT/WIRCam yields parallactic distances for all systems, robust orbit determinations for 23 systems, and photocenter orbits for 19 systems. As a result, we measure 38 precise individual masses spanning 30–115 ${M}_{\mathrm{Jup}}$. We determine a model-independent substellar boundary that is ≈70 ${M}_{\mathrm{Jup}}$ in mass (≈L4 in spectral type), and we validate Baraffe et al. evolutionary model predictions for the lithium-depletion boundary (60 ${M}_{\mathrm{Jup}}$ at field ages). Assuming each binary is coeval, we test models of the substellar mass–luminosity relation and find that in the L/T transition, only the Saumon & Marley "hybrid" models accounting for cloud clearing match our data. We derive a precise, mass-calibrated spectral type–effective temperature relation covering 1100–2800 K. Our masses enable a novel direct determination of the age distribution of field brown dwarfs spanning L4–T5 and 30–70 ${M}_{\mathrm{Jup}}$. We determine a median age of 1.3 Gyr, and our population synthesis modeling indicates our sample is consistent with a constant star formation history modulated by dynamical heating in the Galactic disk. We discover two triple-brown-dwarf systems, the first with directly measured masses and eccentricities. We examine the eccentricity distribution, carefully considering biases and completeness, and find that low-eccentricity orbits are significantly more common among ultracool binaries than solar-type binaries, possibly indicating the early influence of long-lived dissipative gas disks. Overall, this work represents a major advance in the empirical view of very low-mass stars and brown dwarfs.

Export citation and abstract BibTeX RIS

1. Introduction

Mass is the most important property governing the destiny of self-gravitating gaseous objects in hydrostatic equilibrium. Mass determines whether an object becomes a star, generating sufficient energy through nuclear fusion of hydrogen to stabilize itself against gravitational collapse, or a brown dwarf, which primarily supports itself by degeneracy pressure. The dichotomy in the way these two classes of objects satisfy hydrostatic equilibrium results in drastically different evolutionary outcomes. Sufficiently massive objects maintain a relatively stable luminosity and temperature, becoming stars on the main sequence for millions to trillions of years. Meanwhile, objects of substellar mass steadily and drastically decrease in luminosity and temperature. One key observational consequence is that the coldest gaseous objects can be confidently identified as brown dwarfs (e.g., Oppenheimer et al. 1995). However, the masses of most ultracool dwarfs are indeterminate because observables like luminosity and temperature are degenerate between mass and age, such that the mass could be lower or higher if the object is younger or older, respectively.

The evolution of brown dwarfs and the substellar boundary itself are largely understood based on theoretical works over the last two decades (e.g., Saumon et al. 1994; Baraffe et al. 1995; Burrows et al. 1997; Chabrier & Baraffe 1997; Lodders 1999; Chabrier et al. 2000; Burrows et al. 2001; Baraffe et al. 2003; SM08; MacDonald & Mullan 2009; Baraffe et al. 2015). Broadly speaking, the interior physics of evolutionary models over this time has remained the same, while the treatment of the surface boundary conditions has advanced greatly, due to improved molecular line lists, chemistry, and cloud modeling (see a recent review by Marley & Robinson 2015). Clouds in particular have long been thought to be the key to explaining major variations in the emergent flux of brown dwarfs, especially in the change from L to T spectral types (the L/T transition). Understanding clouds in theoretical modeling is imperative, given that they can exert great influence on the atmosphere's opacity, but the physical processes governing cloud formation, particle size distribution, and sedimentation are all effectively free parameters that can only be constrained loosely by observations. Over the past decade, model atmospheres have been largely developed with guidance from observations of colors, absolute magnitudes, and spectra, thus they can reproduce the observed surface properties of ultracool dwarfs reasonably well. However, as atmospheres are a key input to evolutionary models, there arises an entirely different question of whether predictions of fundamental properties (luminosity, radius, etc.) are improved by the adoption of the latest cloud prescriptions and molecular opacities.

Direct mass measurements are central to tests of evolutionary models. Previous work on ultracool dwarfs (spectral types ≥M7) has focused on total dynamical masses that can be readily determined from relative astrometric orbits because the narrow field of view afforded by adaptive optics (AO) imaging rarely contains any reference stars that could be used for absolute astrometry. To date there are total dynamical masses for more than a dozen ultracool binaries (e.g., Lane et al. 2001; Leinert et al. 2001; Bouy et al. 2004; Liu et al. 2008; Cardoso et al. 2009; Dupuy et al. 2009a, 2009b, 2009c, 2010, 2014; Konopacky et al. 2010). In a few special cases, individual masses have been determined: for two single companions to main-sequence stars (Ireland et al. 2008; Crepp et al. 2012); for the companion AB Dor C by assuming a mass for AB Dor A (Close et al. 2005; Guirado et al. 2006); and for three ultracool binary systems where mass ratios are measured using radial velocities or absolute astrometry (Zapatero Osorio et al. 2004; Simon et al. 2006; Seifahrt et al. 2008; Konopacky et al. 2010; Köhler et al. 2012; Dupuy et al. 2016). However, most of the objects in this small sample in the literature are likely to be stars and not brown dwarfs. For additional context, our individual mass measurements for the L6.5+T1.5 binary SDSS J1052+4422AB marked the first individual masses for any field L or T dwarfs (Dupuy et al. 2015b). This stands in stark contrast to the situation for individual dynamical masses of earlier-type M dwarfs, where dozens of measurements over the years have now reached very high precision (e.g., Henry & McCarthy 1993; Delfosse et al. 2000; Benedict et al. 2016). Thus, even though significant progress has been made on ultracool dwarfs using mostly total dynamical masses, the individual masses needed for the strongest tests of models have been lacking.

We present here orbit and mass determinations for a large sample of ultracool binaries with component spectral types of M7–T5. The results here are based on a uniform (re-)analysis of Keck AO imaging and masking data from the past 10 years, as well as Hubble Space Telescope (HST) imaging from the past 20 years, and nearly a decade of absolute astrometric monitoring at the Canada–France–Hawaii Telescope (CFHT). The combined data set not only provides total dynamical masses from the relative orbits and CFHT parallaxes but also yields precise mass ratios from measurements of photocenter motion in our CFHT data. We report observations for our entire sample of 31 binaries, from which we determine robust orbits for 23 systems and individual masses for 19 systems (two of which turn out to be previously unknown triple systems). We critically examine some of the basic predictions of evolutionary models, such as the mass limits for hydrogen and lithium fusion, and we establish empirical relations between parameters such as mass, luminosity, effective temperature, absolute magnitude, and spectral type. We use the brown dwarfs in our sample as clocks (age-dated from models using mass and luminosity) to directly determine the age distribution of the field population. Finally, we revisit the eccentricity distribution of ultracool dwarfs (initially discussed in Dupuy & Liu 2011) and its potential implications for the earliest stages of evolution.

2. Sample Selection for Keck and CFHT Orbit Monitoring

The orbit sample presented here was drawn from the very low-mass visual binaries (${M}_{\mathrm{tot}}\lesssim 0.2$ M; using published mass estimates) known to us in early 2008. In addition to the total mass criterion, we also excluded binaries with integrated-light spectral types of M6 or earlier and published distance estimates ≳40 pc. For the work presented here, we also only consider binaries with types earlier than T6. As the starting point for defining this sample, we used the summary table of Burgasser et al. (2007, pp. 427–441)3 to which we added other binaries from the literature and from our own proprietary Keck LGS AO data in hand at the time. From this whole sample of binaries, we selected only those with estimated periods that suggested ≳30% of their orbit could be complete by 2010. We computed these initial period estimates using the estimated total mass from the literature and a range of semimajor axes corresponding to the projected separation at discovery (e.g., Torres 1999; Dupuy & Liu 2011), where we adopted the 1σ minimum period for our estimates.

In Table 1 we list all 33 binaries in our Keck+CFHT astrometric monitoring sample, along with three other binaries that have published orbit and parallax measurements. We began obtaining resolved Keck AO astrometry in 2007–2008, and we combined our new astrometry with available data in the literature or public archives (e.g., HST and Gemini) to refine our orbital period estimates and thereby our prioritizaton for Keck observations. As described in Dupuy et al. (2011), we performed a Monte Carlo analysis of this multi-epoch data, which provided period estimates based on actual orbital motion, not simply projected separations. We subsequently deprioritized Keck observations of the systems with longer expected periods, focusing our observational efforts on the shorter-period systems. We did occasionally obtain Keck astrometry for the systems with the longest expected periods in our sample.

Table 1.  Orbit Monitoring Sample

Name Binary Discovery Integrated-light Spectral Type Component Spectral Types
  Sep. Epoch (UT) References Optical/IR References A+B References
LP 349-25AB 0farcs125 2004 Jul 03 Forv05 M8/M8 int-g Gizi00, Dupu17 M7:+M8: Dupu12
LP 415-20AB 0farcs119 2002 Feb 07 Sieg03 M7.5/⋯ Gizi00 M6:+M8 Dupu17
SDSSp J042348.57−041403.5AB 0farcs164 2004 Jul 22 Burg05 L7.5/T0 Cruz03, Burg06a L6.5:+T2 Dupu12
2MASS J05185995−2828372AB 0farcs051 2004 Sep 07 Burg06b L7/T1p Kirk08, Burg06a L6:+T4 Dupu12
2MASS J07003664+3157266AB 0farcs170 2004 Dec 29 Reid06a L3.5/⋯ Thor03 L3:+L6.5:. Dupu12
LHS 1901AB 0farcs275 2004 Jan 08 Mont06 M7/M7 Lepi09, Dupu10 M7:+M7: Dupu12
2MASSI J0746425+200032AB 0farcs220 2000 Apr 25 Reid01 L0.5/L1 Kirk00, Knap04 L0+L1.5a Bouy04
2MASSs J0850359+105716AB 0farcs160 2000 Feb 01 Reid01 L6/⋯ Kirk99 L6.5:+L8.5: Dupu12
2MASSI J0856479+223518AB 0farcs098 2001 Apr 24 Bouy03 L3:/⋯ Cruz03
2MASSW J0920122+351742AB 0farcs070 2000 Feb 09 Reid01 L6.5/T0p Kirk00, Burg06a L5.5:+L9:. Dupu12
SDSS J092615.38+584720.9AB 0farcs070 2004 Feb 05 Burg06b ⋯/T4.5 Burg06a T3.5:+T5: Dupu12
2MASSI J1017075+130839AB 0farcs100 2001 Apr 16 Gizi03 L2:/L1 Cruz03, Wils03 L1.5:+L3: Dupu12
SDSS J102109.69−030420.1AB 0farcs172 2004 May 02 Burg06b T3.5/T3 Kirk08, Burg06a T0:+T5 Dupu12
2MASSW J1047138+402649AB 0farcs122 2002 Apr 25 Clos03 M8/⋯ Gizi00a M8+L0: Dupu17
SDSS J105213.51+442255.7AB 0farcs042 2005 May 01 Dupu15 ⋯/T0.5 Chiu06 L6.5:.+T1.5: Dupu15
Gl 417BC 0farcs070 2001 Feb 14 Bouy03 L4.5/⋯ Kirk00 L4.5:+L6: Dupu12
LHS 2397aAB 0farcs270 1997 Apr 12 Free03 M8/⋯ Kirk95 M8a+⋯ Dupu09b
DENIS-P J1228.2−1547AB 0farcs275 1998 Jun 02 Mart99 L5/L6 ± 2 Kirk99, Knap04 L5.5:+L5.5: Dupu12
Kelu-1AB 0farcs291 2005 May 01 Liu05 L2/⋯ Kirk99 L2:+L4: Dupu12
2MASS J14044948−3159330AB 0farcs134 2006 Jun 03 Loop08 T0/T2.5 Loop08, Loop07 L9:+T5 Dupu17
HD 130948BC 0farcs134 2001 Feb 24 Pott02 ⋯/⋯ L4:+L4: Goto02
Gl 569Bab 0farcs101 1999 Aug 28 Mart00 M8.5/⋯ Henr90 M8.5+M9 Lane01
SDSS J153417.05+161546.1AB 0farcs110 2005 May 01 Liu06 ⋯/T3.5 Chiu06 T0:+T5.5 Dupu12
2MASS J15344984–2952274AB 0farcs065 2000 Aug 18 Burg03b T6/T5.5 Burg03a, Burg06a T4.5+T5 Dupu12
2MASSW J1728114+394859AB 0farcs130 2000 Aug 12 Gizi03 L7/⋯ Kirk00 L5:+L7: Dupu12
LSPM J1735+2634AB 0farcs290 2005 Jun 05 Law06 M7.5/⋯ Schm07 M7.5+L0: Dupu12
2MASSW J1750129+442404AB 0farcs158 2002 Apr 25 Sieg03 M7.5/M8 Gizi00b, Dupu12 M6.5:+M8.5: Dupu12
2MASSI J1847034+552243AB 0farcs082 2003 Jul 10 Sieg05 M6.5/⋯ Cruz03 M6+M7 Dupu12
SDSS J205235.31−160929.8AB 0farcs103 2008 Jun 24 Stum11 ⋯/T1: Chiu06 L8.5:.+T1.5 Dupu12
2MASSI J2132114+134158AB 0farcs066 2006 Jun 17 Sieg07 L6/⋯ Cruz07 L4.5:.+L8.5:. Dupu12
2MASSW J2140293+162518AB 0farcs155 2001 Sep 22 Clos02b M8.5/⋯ Gizi00b M8+L0.5: Dupu17
2MASSW J2206228−204705AB 0farcs168 2001 Sep 22 Clos02b M8/M8 Gizi00b, Dupu09a M8+M8 Dupu09a
DENIS-P J225210.7−173013AB 0farcs130 2005 Jun 21 Reid06b ⋯/L7.5 Kend04 L4:+T3.5 Dupu17
Other Ultracool Visual Binaries with Published Orbits and Parallaxes
LHS 1070BC 0farcs267 1993 Jul 29 Lein94 M8.5/⋯ Lein00 M9.5+L0a Rajp12
LSPM J1314+1320AB 0farcs130 2005 Jun 15 Law06 M7/M6 fld-g Lepi09, Dupu16
$\epsilon $ Ind Bab 0farcs732 2003 Aug 13 McCa04 ⋯/T2.5 Scho03 T1+T6 King10

Note.

aThe component spectral types in these cases come from resolved optical spectroscopy. All other component types are based on near-IR spectra, either resolved or deconvolved from the integrated-light near-IR spectrum using near-IR resolved photometry. For LHS 2397aAB, the primary spectral type is assumed to be the same as the optical integrated-light type given the large contrast ratio in the optical.

References. Bouy03—Bouy et al. (2003), Bouy04—Bouy et al. (2004), Burg03a—Burgasser et al. (2003a), Burg03b—Burgasser et al. (2003b), Burg05—Burgasser et al. (2005), Burg06a—Burgasser et al. (2006a), Burg06b—Burgasser et al. (2006b), Chiu06—Chiu et al. (2006), Clos02a—Close et al. (2002a), Clos02b—Close et al. (2002b), Clos03—Close et al. (2003), Cruz03—Cruz et al. (2003), Cruz07—Cruz et al. (2007), Dupu09a—Dupuy et al. (2009a), Dupu09b—Dupuy et al. (2009c), Dupu10—Dupuy et al. (2010), Dupu12—Dupuy & Liu (2012), Dupu15—Dupuy et al. (2015b), Dupu16—Dupuy et al. (2016), Dupu17—this work, Forv05—Forveille et al. (2005), Free03—Freed et al. (2003), Gizi00a—Gizis et al. (2000b), Gizi00b—Gizis et al. (2000a), Gizi03—Gizis et al. (2003), Henr90—Henry & Kirkpatrick (1990), Kend04—Kendall et al. (2004), King10—King et al. (2010), Kirk00—Kirkpatrick et al. (2000), Kirk08—Kirkpatrick et al. (2008), Kirk95—Kirkpatrick et al. (1995), Kirk99—Kirkpatrick et al. (1999), Knap04—Knapp et al. (2004), Lane01—Lane et al. (2001), Law06—Law et al. (2006), Lein00—Leinert et al. (2000), Lein94—Leinert et al. (1994), Lepi09—Lépine et al. (2009), Liu05—Liu & Leggett (2005), Liu06—Liu et al. (2006), Loop07—Looper et al. (2007), Loop08—Looper et al. (2008), Mart00—Martín et al. (2000), Mart06—Martín et al. (2006), Mart99—Martín et al. (1999), McCa04—McCaughrean et al. (2004), Mont06—Montagnier et al. (2006), Phan05—Phan-Bao et al. (2005), Phan06—Phan-Bao et al. (2006), Pott02—Potter et al. (2002), Rajp12—Rajpurohit et al. (2012), Reid01—Reid et al. (2001), Reid06a—Reid et al. (2006a), Reid06b—Reid et al. (2006b), Schm07—Schmidt et al. (2007), Scho03—Scholz et al. (2003), Sieg03—Siegler et al. (2003), Sieg05—Siegler et al. (2005), Sieg07—Siegler et al. (2007), Stum05—Stumpf et al. (2005, p. 8571), Stum11—Stumpf et al. (2011), Thor03—Thorstensen & Kirkpatrick (2003), Wils03—Wilson et al. (2003).

Download table as:  ASCIITypeset image

For most of the binaries in our sample, we began integrated-light astrometric monitoring with CFHT in either the second half of 2007 or in 2008, and continued collecting data until parallaxes were determined. We did not include in our CFHT program the three binaries with Hipparcos parallaxes (Gl 417BC, HD 130948BC, Gl 569Bab), but we did include other systems with published parallaxes at that time (11 binaries) in order to improve the distance accuracy. Indeed, for seven of these binaries, we reduce the uncertainty by a factor of 1.9–6.7. Most of these initial parallax determinations were reported in Dupuy & Liu (2012). We continued monitoring this sample with CFHT up to the present in order to place constraints on photocenter motion and thereby the binary mass ratios.

Our orbit sample selection is essentially based on spectral type and projected separation, but for some purposes, like studies of the eccentricity distribution (e.g., Dupuy & Liu 2011), it is important to consider whether our target selection and prioritization has introduced biases in our sample beyond those intrinsic to the population of ultracool binaries known in late 2007 and early 2008. Although our sample was initially defined in terms of the probability that a given binary would yield an orbit determination by 2010, we can retroactively determine over what range of observable properties our sample is complete. To the best of our knowledge, our initial sample included all observable binaries known at the time with integrated-light spectral types of M6.5–T5.5,4 projected separations ≤6 au at discovery, and distances ≲40 pc (based on the spectrophotometric distance estimate in the absence of a parallax).5 2MASSI J1426316+155701AB is the only binary meeting these criteria for which we never ultimately obtained any CFHT or Keck LGS AO data because we de-prioritized it based on archival and published astrometry (Bouy et al. 2008) that showed its projected separation had increased from 4.1 au at discovery to 6.9 au projected separation by 2006 June. Although we have attempted to resolve 2MASSI J0856479+223518AB on multiple occasions with Keck LGS AO, we have never been successful, so this binary does not appear in any of the following discussion.

Practical observational considerations imposed some limitations on our sample that should not correspond to any physically meaningful selection biases. The largest subset of targets meeting the above criteria but excluded from our sample is that not accessible to Keck AO because the targets are too far south: GJ 1001BC (L5; Golimowski et al. 2004b), DENIS J035726.9−441730AB (L0; Bouy et al. 2003; Gizis et al. 2003), SCR J1845−6357AB (M8.5; Biller et al. 2006), $\epsilon $ Ind Bab (T2.5; McCaughrean et al. 2004), and 2MASS J22551861−5713056AB (L6; Reid et al. 2008a). Another practical limitation is that not all ultracool binaries have a nearby star bright enough for tip-tilt correction as needed for LGS AO imaging at Keck (R ≲ 18.5 mag within ≈65''). This excluded DENIS J100428.3−114648AB (M8), 2MASSW J1239272+551537AB (L5), and 2MASS J14304358+2915405AB (L2). This tip-tilt star limitation also excluded three of the most promising T dwarf binaries, SDSS J0423−0414AB (T0), SDSS J0926+5847AB (T4.5), and 2MASS J0518−2828AB (T1p), so we conducted HST monitoring of these binaries instead.

Some of our sample binaries already have published orbital monitoring results, either from our own work (Liu et al. 2008; Dupuy et al. 2009a, 2009b, 2009c, 2010; Dupuy & Liu 2011; Dupuy et al. 2014, 2015b) or from others (Lane et al. 2001; Zapatero Osorio et al. 2004; Simon et al. 2006; Konopacky et al. 2010). In the following, we perform a complete, uniform analysis of all data on these systems, even in cases where we have previously published results. 2MASS J0746+2000AB had a published orbit (Bouy et al. 2004) before we began our observing program, and Konopacky et al. (2010) provided significant additional astrometry and a refined orbit. We only obtained Keck astrometry for 2MASS J0746+2000AB at one epoch, but we have been monitoring its absolute astrometry from CFHT since 2008. Therefore, we re-analyzed published HST and Keck data of 2MASS J0746+2000AB from public archives in order to include it in our sample of joint Keck+CFHT orbital analysis. LHS 1070BC is the only other binary that would have been in our sample if it did not have a previously published orbit (Leinert et al. 2001; Seifahrt et al. 2008; Köhler et al. 2012). Since the published orbit of LHS 1070BC is not based on Keck astrometry and we did not observe it as part of our CFHT astrometry program, we simply use the orbit parameters quoted in the literature and perform no orbital analysis of our own. We also include the recently published orbit for LSPM J1314+1320AB (Dupuy et al. 2016), which is based on the same orbital analysis presented here. The $\epsilon $ Ind Bab system is too far south to observe from Maunakea, but it also has a published orbit determination from Cardoso et al. (2009), so we include the published parameters in our discussion as well.

3. Observations

3.1. Relative Astrometry from High-angular Resolution Observations

3.1.1. Keck/NIRC2 AO Imaging and Masking

We present here new Keck/NIRC2 AO imaging and non-redundant aperture-masking observations, both in natural guide star and laser guide star modes, in addition to a re-analysis of our own previously published data and publicly available archival data for our sample binaries. Our approach for reducing NIRC2 imaging and obtaining the binary parameters separation, position angle (PA), and flux ratio is well-established from our previous works (Liu et al. 2006, 2008; Dupuy et al. 2009a, 2009b, 2009c, 2010, 2015a). Briefly, we apply standard calibrations (dark subtraction, flat fielding) and then fit an analytic, three-component Gaussian model to each point source in the images. In cases where the binary components are spatially well-separated, we perform PSF-fitting using StarFinder (Diolaiti et al. 2000). In other cases where a third, unsaturated star is in the field (e.g., Gl 569A and Gl 569Bab), we use the third single star as an empirical PSF to fit the binary. Analysis of our NIRC2 masking data was performed using a pipeline similar to previous papers (e.g., Ireland & Kraus 2008; Ireland et al. 2008) and is described in detail in Section 2.2 of Dupuy et al. (2009c). For NIRC2 narrow camera images obtained before 2015 April 13 UT, we use the astrometric solution of Yelda et al. (2010) to correct our measured (x, y) image coordinates for nonlinear optical distortion, using their derived pixel scale of 9.952 ± 0.002 mas pixel−1 and +0fdg252 ± 0fdg0.009 correction for the orientation given in the NIRC2 image headers. For narrow camera imaging obtained after 2015 April 13 UT, we use an updated distortion solution from Service et al. (2016) that accounts for the first-ever realignment of the AO system. The post-realignment pixel scale and orientation are 9.971 ± 0.004 mas pixel−1 and +0fdg262 ± 0fdg0.020, respectively. For NIRC2 wide camera images, which represent a very small subset of our observations, we used the distortion solution of H. Fu et al. (2012, private communication),6 which assumes a pixel scale of 39.686 mas pixel−1 and the Yelda et al. P.A. offset of +0fdg252.

We have made some small improvements to our astrometric analysis compared to our previous work, as described in Dupuy et al. (2016). For data obtained in the vertical angle mode, where the sky rotation of the images relative to the detector axes is constantly changing, we corrected the rotator angles reported in the header to correspond to the midpoint instead of the start of the exposure. We also apply corrections for differential aberration and atmospheric refraction. The refraction correction requires knowledge of the air temperature, pressure, and humidity on Maunakea during our observations, for which we used the weather data archived by CFHT.7 We generally adopt errors that are the rms of the measurements from individual dithers at a given epoch. In some cases where we have previously published astrometry, we instead use estimates of the errors from Monte Carlo simulations of fitting artificial binaries.

Table 2 gives our measured astrometry and flux ratios for all Keck AO data used in our orbital analysis. In total there are 339 distinct measurements (unique bandpass and epoch for a given target), where 302 of these are direct imaging and 37 are non-redundant aperture masking. The median Keck separation error is 0.6 mas with 90% of measurements having errors between 0.06–1.9 mas. The median P.A. error is 0fdg3, with 90% of measurement errors between 0fdg02 and 1fdg3. (We caution that there are possible systematic effects, e.g., uncertainty in the distortion solution and tip-tilt jitter, that are difficult to quantify and may impact the few measurements with astrometric errors well below 0.1 mas. The fact that none of these data points show up as outliers in our orbit fitting analysis described later could be due to the fact that not all phases of all orbits are overconstrained by numerous degrees of freedom.) Eight of the imaging measurements are from six unpublished archival data sets. For HD 130948BC we used data sets from 2005 February 25 UT (PI Prato) and 2011 March 25 UT (PI Bowler). For Gl 569Bab we used data sets from 2003 April 15 UT (PI Simon), 2004 August 10 UT (PI Kulkarni), and 2011 March 25 UT (PI Bowler). For LSPM J1735+2634AB we used a JHK data set from 2007 August 1 UT (PI N. Law). Fourteen other measurements are from our re-analysis of 10 previously published data sets in the archive: LP 349-25, 2MASS J0920+3517AB, 2MASS J0746+2000AB, HD 130948BC, 2MASS J1847+5522AB, and 2MASS J2206−2047AB (PI Ghez; Konopacky et al. 2010) and SDSS J2052−1609AB (PI C. Gelino; Bardalez Gagliuffi et al. 2015).

Table 2.  Relative Astrometry from Keck/NIRC2 Adaptive Optics Imaging and Masking

Observation Date Separation PA Δm Bandpass Nframes Notes
(UT) (MJD) (mas) (°) (mag)      
LP 349-25AB (Nep = 8, Δt = 3.99 years)
2008 Jan 16 54481.25 137.25 ± 0.24 208.07 ± 0.06 0.335 ± 0.016 ${K}_{S}$ 10 I
2008 Jun 30 54647.55 114.88 ± 0.25 194.94 ± 0.12 0.285 ± 0.011 ${K}_{S}$ 9 I
2008 Jun 30 54647.56 114.67 ± 0.23 195.13 ± 0.08 0.303 ± 0.004 H 7 I
2008 Jun 30 54647.56 114.5 ± 0.6 195.20 ± 0.17 0.345 ± 0.011 J 11 I
2008 Aug 20 54698.61 105.71 ± 0.10 189.90 ± 0.05 0.353 ± 0.010 ${K}_{S}$ 12 I
2008 Sep 9 54718.58 102.17 ± 0.14 187.62 ± 0.09 0.243 ± 0.005 L' 12 I
2009 Sep 28 55102.56 71.20 ± 0.30 98.4 ± 0.6 0.230 ± 0.030 ${K}_{S}$ 14 I
2009 Dec 15 55180.37 83.6 ± 0.4 81.40 ± 0.30 0.37 ± 0.06 K 14 I
2010 May 22 55338.62 112.76 ± 0.18 59.93 ± 0.09 0.397 ± 0.019 K 13 Ia
2012 Jan 14 55940.25 131.4 ± 0.7 19.52 ± 0.12 0.258 ± 0.011 K 15 I

Notes. In the notes column, "I" indicates an observation done with direct imaging and "M" indicates non-redundant aperture masking.

aThis denotes a previously published data set that we obtained from the public NIRC2 archive. bThis denotes a previously unpublished data set that we obtained from the public NIRC2 archive.

Only a portion of this table is shown here to demonstrate its form and content. A machine-readable version of the full table is available.

Download table as:  DataTypeset image

3.1.2. HST/ACS-WFC Monitoring

In addition to our Keck AO monitoring, we also obtained data for three T dwarf binaries lacking suitable LGS tip-tilt stars over a three-year HST program8 using the Advanced Camera for Surveys (ACS) Wide Field Camera (WFC). ACS-WFC has a pixel scale of ≈50 mas, and we used the F814W bandpass for all of our observations to optimize between S/N and angular resolution. SDSS J0423−0414AB (T0) and SDSS J0926+5847AB (T4.5) were both resolved in all of our images. 2MASS J0518−2828AB (T1p) was not resolved in any of our six observations spanning 2009 December 11 UT to 2012 October 27 UT. Based on simulations we performed for our observation proposal, we expected to readily resolve 2MASS J0518−2828AB if it was at a similar separation as its discovery (51 mas; Burgasser et al. 2006b). Given that the flux ratio of 2MASS J0518−2828AB in the F814W band is unknown (and difficult to extrapolate from the available infrared data), and a larger flux ratio of ≈2 mag could make it quite difficult to clearly resolve, we conservatively estimate an upper limit of <50 mas on the putative companion's separation during our observations.

For the resolved binaries, we performed PSF fitting of our ACS-WFC images in a similar manner to our previous work on data from other HST cameras (e.g., Liu et al. 2008; Dupuy et al. 2009a, 2009b). We used TinyTim (Krist et al. 2011) to generate a PSF model for each pipeline-reduced (flt) FITS image that we downloaded from the HST archive. The only inputs to this model are the target's position on the detector and a template spectrum of comparable spectral type to the target. For the latter, we used optical-to-infrared spectra from Geballe et al. (2002), adopting the integrated-light spectrum of SDSS J0423−0414AB9 for each of its components and the spectrum of 2MASS J0559−140410 for the components of SDSS J0926+5847AB. To test the impact of assuming the same spectrum for each component, we also tried L5 and T5 templates for both components of SDSS J0423−0414AB and found the astrometry changed by <0.6 mas (<1σ). We use the same PSF model for both components and shift, scale, and add them to each other to create a model binary image. We used the amoeba function in IDL, based on the routine from Press et al. (1992), to find the best-fit parameters for the binary model: position and flux of the primary (x0, y0, f0) and the binary separation, P.A., and flux ratio. We used the information contained in the FITS files to correct our fitted (x, y) positions on the detector to sky coordinates. This accounts for detector defects from the D2IMARR FITS extension, a polynomial distortion correction contained in FITS header keywords, and non-polynomial distortion corrections from the WCSDVARR FITS extension (Kozhurina-Platais et al. 2015). We then applied the CD matrix in the FITS header and a tangent projection to convert the corrected x and y values to R.A. and decl.

Table 3 reports the mean and rms of our best-fit binary parameters derived from the individual dithered ACS-WFC images for SDSS J0926+5847AB. Although the first five observations show the binary moving outward and then inward slightly (67 mas to 73 mas to 63 mas), at the final epoch it had moved in significantly and was only marginally resolved. Over the first five epochs we measured flux ratios consistent with being constant at ${\rm{\Delta }}F814W=0.560\pm 0.024$ mag. We therefore fixed the flux ratio to be 0.56 mag in our fitting of the final imaging observations, giving a separation of 33 ± 5 mas and a very uncertain P.A. of 283° ± 30°. We note that there is nominally an ambiguity in this measured P.A., typical for marginally resolved binaries, such that it could be 103° ± 30° instead. However, 283° ± 30° is in better agreement with the orbit fit, so we adopt this value.

Table 3.  Relative Astrometry from HST Imaging and the Literature

Date Separation P.A. Δm Bandpass Notes
(UT) (mas) (°) (mag)    
LP 349-25AB (Nep = 2)
2004 Jul 3 125 ± 10 12.7 ± 2.0 0.26 ± 0.05 K' CFHT/PUEO (Forveille et al. 2005)
2004 Sep 26 107 ± 10 7.1 ± 0.5 0.38 ± 0.05 H VLT/NACO (Forveille et al. 2005)
LP 415-20AB (Nep = 1)
2002 Feb 7 119.0 ± 1.0 91.2 ± 0.7 0.66 ± 0.06 K' Gemini/QUIRC (Siegler et al. 2003)

Note. The measurements reported in this table are from our own analysis of archival data in this paper, unless another reference is given in the Notes column.

Only a portion of this table is shown here to demonstrate its form and content. A machine-readable version of the full table is available.

Download table as:  DataTypeset image

At separations of 104–151 mas, SDSS J0423−0414AB was much wider in our ACS-WFC imaging than SDSS J0926+5847AB. We found that using the rms of our SDSS J0423−0414AB measurements as the uncertainties produced a somewhat high χ2 for the final orbit fit, implying that the rms does not fully capture all systematic errors for this well-resolved binary. We therefore examined the rms about the orbit fit of our ACS-WFC data and used this as the common uncertainty in separation (0.6 mas) and P.A. (0fdg7) for all epochs of SDSS J0423−0414AB imaging. The F814W flux ratio we measured across multiple epochs shows some variation within the rms errors, but it would be consistent with being constant if there are 0.05 mag systematic errors in our measured SDSS J0423−0414AB flux ratios.

3.1.3. Other Published and Archival HST Imaging

Many of our sample binaries have HST imaging data in the public archive. These are often the observations that discovered the binaries initially, providing the first epoch of astrometry that we use in our orbit fits. As in our previous work, we have re-analyzed the available archival data using our own TinyTim PSF-fitting routine as described in Section 3.1.2. These data come from the WFPC2 Planetary Camera (WFPC2-PC1), ACS High Resolution Channel (ACS-HRC), and NICMOS Camera 1 (NICMOS-NIC1). For all three of these cameras, numerous images of single ultracool dwarfs are also available in the archive from various imaging surveys. We collected these images to make a library of observed single PSFs for each relevant camera and filter combination and used them to perform Monte Carlo simulations of our PSF fitting. For a given binary observation, we replicated the binary properties as closely as possible to the actual fractional pixel separation in x and y by co-adding two different library PSFs centered at different subpixel locations, and scaling the two PSFs to match the flux ratio and S/N of the actual binary data. We ran 100 simulations for each observation, applied the mean systematic offset as a correction to our measurement, and used the rms as the error.

Table 3 gives the resulting binary parameters from our analysis of archival WFPC2-PC1, ACS-HRC, and NICMOS-NIC1 imaging. WFPC2-PC1 data come from programs GO-6345 (PI Kirkpatrick), GO-8146 (PI Reid), GO-8581 (PI Reid), GO-8563 (PI Kirkpatrick), GO-8720 (PI Brandner), GO-9157 (PI Martín), GO-9345 (PI Martín), and GO-9968 (PI Martín). ACS-HRC data come from programs GO-9451 (PI Brandner) and GO-10559 (PI Bouy). NICMOS-NIC1 data come from programs GO-7952 (PI Martín), GO-9833 (PI Burgasser), GO-9843 (PI Gizis), GO-10143 (PI Reid), and GO-11136 (PI Liu).

3.2. Absolute Astrometry from CFHT/WIRCam

We obtained high-precision unresolved astrometry of our sample to determine parallaxes and proper motions, and to constrain the photocenter motion due to the binary orbits. We present here an updated analysis of our data from the Hawaii Infrared Parallax Program that uses the CFHT facility infrared camera WIRCam (Puget et al. 2004). Our observing strategy and custom astrometry pipeline are described in detail in Dupuy & Liu (2012). Briefly, for a given target we obtain data sets on multiple nights each season, with each data set typically comprising 20–30 dithered frames in the J band, or in a narrow K-band filter (${K}_{{\rm{H}}2}$) if the target would saturate in J. Queue observing constraints at CFHT are used to require good seeing and that all J-band observations occur near transit, i.e., at similar airmass, to guard against systematic astrometric errors due to differential chromatic refraction (DCR). We measure Gaussian-windowed centroids by running Sextractor v2.19.5 (Bertin & Arnouts 1996) on the detrended images provided by CFHT. We cross-match detections at a given epoch, using our custom distortion solution and simple linear transformations, and adopt the standard error on the mean as the astrometric measurement error for any given star at each epoch. We then solve for the linear transformations between epochs, iteratively solving for and masking high proper motion and/or parallax sources. We finally cross-match to an astrometric reference catalog (2MASS or SDSS-DR9) to provide absolute calibration of the linear terms and then solve for the proper motion and parallax of every star in the field.

Unlike our previous work, we do not use a stand-alone analysis of our CFHT data to compute the parallax and proper motion of our target binaries. This approach was used by Dupuy & Liu (2012), where we then applied evolutionary model-based corrections to account for the expected orbital motion in the CFHT photocenter. As we have continued monitoring our sample binaries from 2007 or 2008 up to the present, we now have data spanning a much longer fraction of many of our targets' orbital periods. It therefore becomes even more important to freely fit for photocenter motion (as described in Section 4), without imposing assumptions about mass–magnitude relations from evolutionary models or potentially unresolved multiplicity. Indeed, the fact that we are now in a position to detect significant photocenter motion allows us to place strong empirical constraints on the relationship between mass and magnitude, as we have recently demonstrated with SDSS J1052+4422AB (Dupuy et al. 2015b), and possibly detect higher-order multiplicity. We have also made some small changes to improve the performance of our pipeline compared to Dupuy & Liu (2012), such as using the mean instead of the median measured position of a given source at each epoch. In addition, when computing the correction from relative to absolute parallax and proper motion using the Besançon model of the Galaxy (Robin et al. 2003), we now exclude model stars with proper motions larger than the cutoff value we use in our iterative solution of the epoch-to-epoch linear terms.

Table 4 gives all of the absolute astrometry from CFHT for the sample of binaries for which we have orbit-monitoring data from Keck and HST. Many of these CFHT observations were originally published in Dupuy & Liu (2012), but the data presented here are nonetheless distinct from what was published previously since we no longer attempt to remove any photocenter motion from the reported absolute astrometry.

Table 4.  Integrated-light Astrometry from CFHT/WIRCam

Observation Date R.A. Decl. ${\sigma }_{{\rm{R}}.{\rm{A}}.}\cos \delta $ σDecl. Airmass Seeing
(UT) (MJD) (deg) (deg) (mas) (mas)   (arcsec)
LP 349-25AB (Nep = 20, Δt = 8.26 years)
2008 Aug 9 54687.5662 006.98451510 +22.32547171 2.4 2.6 1.001 0.61
2008 Sep 6 54715.4857 006.98451751 +22.32546921 3.0 8.6 1.001 0.48
2008 Oct 8 54747.4233 006.98451811 +22.32545820 2.8 3.3 1.008 0.53
2008 Oct 9 54748.3920 006.98451801 +22.32545770 2.7 3.5 1.002 0.75
2008 Nov 16 54786.2900 006.98451972 +22.32544801 1.9 3.2 1.002 0.90
2009 Jul 29 55041.6143 006.98463346 +22.32542572 2.6 4.0 1.006 0.84
2009 Aug 4 55047.6397 006.98463611 +22.32542544 3.1 3.6 1.063 0.56
2009 Aug 8 55051.5876 006.98463564 +22.32542301 2.0 2.3 1.006 0.68
2009 Aug 25 55068.5222 006.98463634 +22.32542258 2.8 2.1 1.001 0.68
2009 Oct 21 55125.3717 006.98463934 +22.32540909 2.8 2.5 1.001 0.71
2009 Nov 6 55141.3542 006.98463957 +22.32540230 8.4 8.7 1.016 0.79
2009 Dec 22 55187.2217 006.98464710 +22.32538996 1.6 1.8 1.010 0.58
2010 Aug 15 55423.5592 006.98475726 +22.32537588 1.4 2.4 1.002 0.58
2010 Sep 15 55454.4572 006.98476077 +22.32536973 3.1 2.8 1.003 0.47
2010 Oct 22 55491.3577 006.98475905 +22.32536073 5.2 3.8 1.002 0.63
2011 Jul 19 55761.6019 006.98487661 +22.32533064 3.3 1.5 1.012 0.53
2011 Jul 25 55767.5641 006.98487564 +22.32533050 3.1 4.2 1.039 0.69
2016 Aug 17 57617.5512 006.98549184 +22.32509917 4.6 5.2 1.002 0.46
2016 Sep 10 57641.4343 006.98549407 +22.32509416 3.6 5.6 1.038 0.57
2016 Nov 13 57705.2584 006.98549810 +22.32507409 18.4 12.0 1.040 0.51

Only a portion of this table is shown here to demonstrate its form and content. A machine-readable version of the full table is available.

Download table as:  DataTypeset image

4. Joint Orbit and Parallax Analysis

We performed a Markov Chain Monte Carlo (MCMC) analysis to simultaneously fit a relative orbital solution and an absolute astrometric solution to our resolved (Keck, HST, etc.) and unresolved (CFHT) astrometry, respectively. This joint fitting approach was originally motivated by binaries that show significant photocenter motion in our CFHT data. However, this approach also has the advantage of appropriately marginalizing over the uncertainty in CFHT photocenter motion when determining parallaxes and proper motions. Fitting for photocenter motion, even if it ends up being consistent with zero, also allows us to constrain mass ratios given that we have independently measured flux ratios. For example, most binaries in our sample have flux ratios near unity, so if their mass ratios are not near unity due to unresolved triple components, then they would display photocenter motion commensurate with the amount of unresolved mass present.

Our joint orbit and parallax MCMC analysis method is similar to our previous work on SDSS J1052+4422AB (Dupuy et al. 2015b). Six orbital parameters are shared in common between the resolved and absolute astrometric solutions: period (P), eccentricity (e), inclination (i), P.A. of the ascending node (Ω), argument of periastron (ω), and mean longitude at the reference epoch (λref).11 The reference epoch (tref) is defined to be 2010 January 1 00:00 UT (2455197.5 JD) and is related to the time of periastron passage, ${T}_{0}={t}_{\mathrm{ref}}-P\tfrac{{\lambda }_{\mathrm{ref}}-\omega }{360^\circ }$. For the resolved orbit we fit for the total semimajor axis ($a={a}_{1}+{a}_{2}$, where a1 and a2 are the semimajor axes of the primary and secondary components about the system barycenter). For the unresolved orbit we fit for a photocenter semimajor axis ${a}_{\mathrm{phot}}$. For the absolute astrometry we also fit for the usual five parameters of parallax (π), position in R.A. (α) and decl. (δ) at a reference epoch, which we choose to be the same as tref above, and proper motion in R.A. (${\mu }_{{\alpha }^{* }}\equiv {\mu }_{\alpha }\cos \delta $) and decl. (${\mu }_{\delta }$). For a circular orbit the parameter ω has no physical meaning. Therefore, in order to allow our MCMC to robustly explore parameter space for nearly circular orbits we chose to step in $\omega +{\rm{\Omega }}$ and $\omega -{\rm{\Omega }}$. Our priors are uniform in e, ω, Ω, λref, and the ratio of ${a}_{\mathrm{phot}}$/a. We adopt log-flat priors in P and a by multiplying likelihoods by 1/P and 1/a. We allow for inclinations to be randomly distributed in space by multiplying likelihoods by $\sin (i)$. Our prior on the parallax assumes a uniform density in space volume, and thus we multiply our likelihood by 1/π2.

To determine the starting point of our chains, we made an initial estimate of the best-fit parameters. For the relative orbit parameters, we performed a grid search over the parameters P, e, and T0. Once these three parameters are specified, then the eccentric anomaly can be computed and the best-fit orbit can be directly analytically determined, as described in detail by Lucy (2014). We searched across 104 randomly drawn values between 0 ≤ e < 1, 0 ≤ T0/P < 1, and $\mathrm{log}(P)$ initially from 103 to 106 days. At subsequent iterations we refined the range of P to center on the best-fit orbital period from the previous trials. After three iterations, we passed the single set of orbit parameters with the lowest χ2 to our custom least-squares minimization routine for orbit fitting (Dupuy et al. 2010), which is based on the MPFIT package in IDL (Markwardt 2009), in order to optimize our starting position for the MCMC. For the parallax parameters, including the photocenter orbit size ${a}_{\mathrm{phot}}$, we used the same method as described in Section 2.4.1 of Dupuy & Liu (2012) to find the best-fit values. Briefly, we used our best-fit orbit parameters, the binary flux ratio for the CFHT bandpass (either the J or K band), and an estimated mass ratio from Cond evolutionary models (Baraffe et al. 2003). Combining the flux ratio and mass ratio gives the scale factor by which to multiply a in order to derive ${a}_{\mathrm{phot}}$. We then subtract the estimated photocenter orbit offsets from the CFHT astrometry and find the best-fit parallax, proper motion, and R.A. and decl.

We used the Python implementation of the Goodman & Weare (2010) affine-invariant ensemble sampler emceev2.1.0 (Foreman-Mackey et al. 2013) to perform our MCMC analysis. We initialized 103 walkers with our best-fit parameters after adding a small amount of Gaussian noise having standard deviations of P × 10−4 in P; a × 10−4 in a; 10−4 in e; 10−4 radians in i, $\omega -{\rm{\Omega }}$, $\omega +{\rm{\Omega }}$, and λref; 10−6 degrees in R.A. and decl.; 10−8 degrees yr−1 in R.A. and decl. proper motion; π × 10−3 in π; and 10−3 in ${a}_{\mathrm{phot}}$/a. We allowed our 103 walkers to run for 105 steps, saving results every 500 steps. After removing the first 50% of the chains as burn-in, our final chains possess a total of 1.0 × 105 values for each parameter.

Our CFHT astrometry gives a measurement of the parallax and proper motion relative to the grid of reference stars. In order to determine distances and absolute proper motions, we estimate the parallax and proper motion of our reference grids using the Besançon model of the Galaxy (Robin et al. 2003)12 in a similar fashion to our previous work (Dupuy & Liu 2012; Dupuy et al. 2015b). The only change we have adopted here is to exclude Besançon model stars with large proper motions and parallaxes. This accounts for the fact that our astrometric pipeline always excludes such stars from our reference grid, using cuts of 10 mas in parallax and 30 mas yr−1 in proper motion. Therefore, our relative-to-absolute corrections reported here are slightly smaller in amplitude than our past work. For example, in Dupuy et al. (2015b) we derived corrections of Δπ = 1.7 ± 0.3 mas, ${\rm{\Delta }}{\mu }_{{\alpha }^{* }}=-6\pm 3$ mas yr−1, and ${\rm{\Delta }}{\mu }_{\delta }=-7\pm 3$ mas yr−1 for SDSS J1052+4422AB, but here we derive ${1.33}_{-0.16}^{+0.13}$ mas, −3.0 ± 1.4 mas yr−1, and −4.2 ± 1.2 mas yr−1, respectively. This results in a <0.5σ change in our absolute parallax for SDSS J1052+4422AB (38.4 ± 0.7 mas previously, 38.1 ± 0.6 mas here).

Table 5 presents for each binary the median, best fit, and 1σ and 2σ credible intervals for the 13 parameters in our joint orbit and parallax analysis. These tables also report various properties of interest that can be derived from our fitted parameters, such as T0, distance (d), and semimajor axis in au. We also report the total system mass for each binary, ${M}_{\mathrm{tot}}/{M}_{\mathrm{Jup}}=1047.93\times {(a/\mathrm{au})}^{3}{(P/\mathrm{yr})}^{-2}$, both with and without including the uncertainty in the parallax so that future improvements in distances for our sample binaries can be easily adapted into improved masses. Table 6 gives the χ2 values of the best-fit solutions as well as the final mean acceptance fraction for each MCMC chain. Figure 1 shows all of the relative and absolute astrometry used in our joint orbit/parallax analysis for each binary, along with the resulting astrometric solution, and Figure 2 shows the resulting posterior distributions from our MCMC analysis.

Figure 1.

Figure 1.

Top left: resolved relative astrometry (filled symbols) shown alongside the best-fit orbit (thick black line) and 100 randomly drawn orbits from our MCMC chain (thin gray lines). The plotting symbols are typically larger than the error bars. Open blue circles indicate the epochs at which we obtained unresolved CFHT/WIRCam astrometry. Top right: our relative astrometry shown as a function of time (top subpanels) and after subtracting the best-fit orbit solution (bottom subpanels). Bottom: CFHT/WIRCam astrometry from unresolved, seeing-limited imaging. Top panels show the data with the proper motion and photocenter orbital motion subtracted in order to display our best-fit parallax solution (thick black line). Bottom panels show the residuals and our best-fit astrometric solution (thick black line) after subtracting our best-fit parallax and linear motion as a function of time (including both true proper motion and linear orbital motion). The figure set contains all the figures. (The complete figure set (31 images) is available.)

Standard image High-resolution image
Figure 2.

Figure 2.

Posterior distributions of orbital parameters from our MCMC analysis. In histograms, solid lines show the medians, dotted lines show the best-fit values, and dashed lines show the 68.3% (1σ) credible intervals. In contour plots, regions containing 68.3% and 95.4% of the posterior are indicated by black dashed lines and gray dashed–dotted lines, respectively. Top right: the final posterior in total system mass after including both the orbit and parallax uncertainties. Middle triangle plot: histograms and correlations for the astrophysical parameters of mass, period, semimajor axis, and eccentricity. In the triangle plot neither mass nor semimajor axis includes the error in distance from the parallax, thus showing the uncertainties and correlations in parameters from the orbit fit alone. Bottom: histograms of the various viewing angles that are part of the orbit fit. (The complete figure set (31 images) is available.)

Standard image High-resolution image

Table 5.  MCMC Posteriors for the Orbit and Parallax

Property Median ± 1σ Best Fit 95.4% C.I. Prior/Notes
Fitted Parameters for LP 349-25AB
Orbital period P (years) 7.698 ± 0.014 7.701 7.671, 7.727 1/P (log-flat)
Semimajor axis a (mas) ${145.99}_{-0.18}^{+0.17}$ 145.98 145.65, 146.33 1/a (log-flat)
Eccentricity e ${0.0468}_{-0.0018}^{+0.0019}$ 0.0470 0.0431, 0.0505 uniform, 0 ≤ e < 1
Inclination i (°) ${117.36}_{-0.10}^{+0.11}$ 117.36 117.16, 117.57 $\sin (i)$, 0° < i < 180°
PA of the ascending node Ω (°) 36.64 ± 0.10 36.62 36.43, 36.84 uniform
Argument of periastron ω (°) 262.2 ± 1.8 262.2 258.6, 266.0 uniform
Mean longitude at 2455197.5 JD λref (°) ${294.44}_{-0.22}^{+0.23}$ 294.40 293.98, 294.89 uniform
${\rm{R}}.{\rm{A}}.-{\rm{R}}.{\rm{A}}{.}_{2010}$ (mas) $-{0.1}_{-1.3}^{+1.2}$ 0.0 −2.6, 2.4 uniform, ${\rm{R}}.{\rm{A}}{.}_{2010}=6.9846505$
$\mathrm{decl}.-\mathrm{decl}{.}_{2010}$ (mas) 0.0 ± 0.9 0.0 −1.8, 1.8 uniform, decl.2010 = +22.3253882
Relative proper motion in R.A. ${\mu }_{{\rm{R}}.{\rm{A}}.,\mathrm{rel}}$ (mas yr−1) 406.4 ± 0.4 406.4 405.7, 407.1 uniform
Relative proper motion in decl. ${\mu }_{\mathrm{decl}.,\mathrm{rel}}$ (mas yr−1) −167.1 ± 0.5 −166.8 −168.0, −166.2 uniform
Relative parallax πrel (mas) 68.0 ± 0.9 68.1 66.3, 69.8 1/π2 (uniform volume density)
Ratio of photocenter orbit to semimajor axis α/a ${0.056}_{-0.008}^{+0.007}$ 0.057 0.041, 0.071 uniform
Derived Properties for LP 349-25AB
Total mass at fixed distance ${a}^{3}{P}^{-2}{(d/14.45\mathrm{pc})}^{3}$ (${M}_{\mathrm{Jup}}$) 165.8 ± 0.7 165.7 164.5, 167.1
Time of periastron T0 (JD) ${2457758}_{-14}^{+15}$ 2457758 2457729, 2457787
Photocenter semimajor axis α (mas) ${8.2}_{-1.2}^{+1.1}$ 8.3 6.0, 10.4
Correction to absolute R.A. proper motion ΔμR.A. (mas yr−1) 1.5 ± 1.6 1.5 −1.8, 4.7
Correction to absolute decl. proper motion Δμdecl. (mas yr−1) −3.4 ± 1.2 −3.4 −5.8, −1.0
Correction to absolute parallax Δπ(mas) ${1.21}_{-0.17}^{+0.14}$ 1.21 0.92, 1.55
Absolute proper motion in R.A. ${\mu }_{{\rm{R}}.{\rm{A}}.,\mathrm{abs}}$ (mas yr−1) 407.9 ± 1.7 407.9 404.5, 411.2
Absolute proper motion in decl. ${\mu }_{\mathrm{decl}.,\mathrm{abs}}$ (mas yr−1) −170.4 ± 1.3 −170.2 −173.0, −167.8
Absolute parallax πabs (mas) 69.2 ± 0.9 69.3 67.5, 71.0
Distance d (pc) ${14.45}_{-0.19}^{+0.18}$ 14.42 14.08, 14.82
Semimajor axis a (au) 2.109 ± 0.027 2.105 2.054, 2.163
Total mass ${M}_{\mathrm{tot}}$ (${M}_{\mathrm{Jup}}$) ${166}_{-7}^{+6}$ 165 154, 179

Only a portion of this table is shown here to demonstrate its form and content. A machine-readable version of the full table is available.

Download table as:  DataTypeset image

Table 6.  Summary of MCMC Results

  13 Parameter Fit Relative Orbit Acceptance
Name χ2 dof χ2 dof Fraction
LP 349-25AB 45.1 51 12.6 17 0.369
LP 415-20AB 50.9 69 20.2 17 0.292
SDSS J0423−0414AB 53.5 59 12.1 15 0.369
2MASS J0700+3157AB 70.1 65 35.6 33 0.132
LHS 1901AB 48.5 47 24.8 23 0.369
2MASS J0746+2000AB 42.2 41 8.9 13 0.369
2MASS J0850+1057AB 51.0 43 16.7 15 0.068
2MASS J0920+3517AB 90.2 89 42.1 43 0.368
SDSS J0926+5847AB 29.9 29 7.7 7 0.170
2MASS J1017+1308AB 50.4 49 11.5 11 0.355
SDSS J1021−0304AB 27.9 25 9.1 7 0.056
2MASS J1047+4026AB 74.8 61 17.4 15 0.369
SDSS J1052+4422AB 66.2 67 17.0 23 0.369
Gl 417BC 21.6 17 0.485
LHS 2397aAB 74.6 79 25.4 23 0.369
DENIS J1228−1557AB 38.6 47 14.8 27 0.093
Kelu-1AB 39.1 43 13.9 25 0.321
2MASS J1404−3159AB 67.0 55 13.5 13 0.090
HD 130948BC 36.9 31 0.486
Gl 569Bab 37.1 31 0.486
SDSS J1534+1615AB 30.3 31 5.0 5 0.072
2MASS J1534−2952AB 66.6 63 34.6 31 0.170
2MASS J1728+3948AB 50.0 63 16.8 33 0.068
LSPM J1735+2634AB 62.7 57 18.4 21 0.367
2MASS J1750+4424AB 31.1 31 6.0 5 0.053
2MASS J1847+5522AB 45.4 39 15.6 13 0.056
SDSS J2052−1609AB 59.3 61 19.9 19 0.087
2MASS J2132+1341AB 75.9 75 18.4 21 0.364
2MASS J2140+1625AB 63.5 65 33.4 27 0.368
2MASS J2206−2047AB 51.8 61 14.1 13 0.230
DENIS J2252−1730AB 51.4 57 18.4 21 0.369

Note. The minimum χ2 and degrees of freedom (dof) for the full 13 parameter MCMC fits are given for all objects with absolute astrometry from CFHT/WIRCam. We also report the χ2 of the best-fit solution computed using only the relative orbit data (Keck, HST, etc.). The last column gives the mean final acceptance fraction for the MCMC analysis as reported by emcee, averaging over all parameters and walkers.

Download table as:  ASCIITypeset image

4.1. Assessing Reliability of Orbit Determinations

Although we have applied our MCMC analysis to all of our astrometric data, not all binaries will have orbit determinations sufficiently reliable for astrophysical use. The primary quality metric for assessing our orbit fits should be the precision with which they constrain the system mass, as very large mass uncertainties will be of little use for constraining models in the following analysis. In addition, if orbital parameters are poorly determined, then the resulting posteriors will be strongly influenced by our adopted priors. Although we have attempted to choose priors that are as uninformed as possible, we should not rely on them to constrain physical parameters (P, a, e) that are not sufficiently constrained by our data. Finally, some of our orbit determinations indicate that our observations span only a small fraction of the full orbital period. Therefore, we also consider this as a possible indicator of whether our data has provided reliable constraints on orbital parameters.

Table 7 provides a summary of the orbit quality metrics that we use to assess the reliability of our orbit fits. The metrics $\delta \mathrm{log}{M}_{\mathrm{tot}}$ and δe are computed as the difference between the maximum and minimum credible interval (1σ) of the total mass at fixed distance and eccentricity, respectively, and the metric ${\rm{\Delta }}{t}_{\mathrm{obs}}/P$ is the fraction of the orbital period (median of the posterior) covered by our resolved astrometry. We now consider our orbit fits, beginning with the apparently least reliable, and moving toward binaries with better reliability metrics until we finally determine which orbits to carry forward in our analysis.

Table 7.  Summary of Orbit Quality Metrics

    Orbit Quality Metrics    
Name SpT ${\rm{\Delta }}\mathrm{log}{M}_{\mathrm{tot}}$ Δe Δtobs/P P ${M}_{\mathrm{tot}}$
  (A+B) (dex)     (years) (${M}_{\mathrm{Jup}}$)
Orbit Determinations Used in Our Analysis
LP 349-25AB M7+M8 0.0035 0.0037 0.978 7.698 ± 0.014 ${166}_{-7}^{+6}$
LP 415-20AB M6+M8 0.0586 0.0233 0.673 14.82 ± 0.24 ${248}_{-29}^{+26}$
SDSS J0423−0414AB L6.5+T2 0.0093 0.0147 0.659 12.30 ± 0.06 83 ± 3
2MASS J0700+3157AB L3+L6.5 0.0035 0.0128 0.431 23.9 ± 0.5 ${141}_{-5}^{+4}$
LHS 1901AB M7+M7 0.0029 0.0017 0.497 16.21 ± 0.08 ${213}_{-10}^{+9}$
2MASS J0746+2000AB L0+L1.5 0.0050 0.0007 0.760 ${12.736}_{-0.029}^{+0.031}$ 163 ± 5
2MASS J0920+3517AB L5.5+L9 0.0098 0.0130 1.822 7.258 ± 0.009 187 ± 11
2MASS J1017+1308AB L1.5+L3 0.0119 0.0196 0.647 ${18.60}_{-0.23}^{+0.22}$ ${156}_{-18}^{+14}$
2MASS J1047+4026AB M8+L0 0.0048 0.0027 2.136 ${6.562}_{-0.026}^{+0.029}$ ${178}_{-12}^{+11}$
SDSS J1052+4422AB L6.5+T1.5 0.0096 0.0046 1.048 ${8.608}_{-0.024}^{+0.025}$ ${90}_{-5}^{+4}$
Gl 417BC L4.5+L6 0.0114 0.0064 0.846 15.65 ± 0.08 ${99.2}_{-3.3}^{+3.0}$
LHS 2397aAB M8+... 0.0072 0.0063 1.178 14.37 ± 0.05 ${159}_{-8}^{+7}$
Kelu-1AB L2+L4 0.0053 0.0098 0.667 24.98 ± 0.19 ${180}_{-26}^{+22}$
2MASS J1404−3159AB L9+T5 0.0445 0.0106 0.536 ${16.52}_{-0.22}^{+0.21}$ ${120}_{-13}^{+11}$
HD 130948BC L4+L4 0.0023 0.0034 1.063 10.009 ± 0.010 ${115.4}_{-2.1}^{+2.2}$
Gl 569Bab M8.5+M9 0.0040 0.0020 5.006 2.3707 ± 0.0005 138 ± 7
2MASS J1534−2952AB T4.5+T5 0.0060 0.0055 0.719 ${20.35}_{-0.06}^{+0.05}$ 99 ± 5
2MASS J1728+3948AB L5+L7 0.0089 0.0277 0.337 ${40.8}_{-1.2}^{+1.6}$ ${140}_{-8}^{+7}$
LSPM J1735+2634AB M7.5+L0 0.0017 0.0075 0.286 ${21.65}_{-0.23}^{+0.24}$ 187 ± 7
2MASS J2132+1341AB L4.5+L8.5 0.0192 0.0094 0.576 ${10.74}_{-0.17}^{+0.16}$ ${128}_{-8}^{+7}$
2MASS J2140+1625AB M8+L0.5 0.0101 0.0146 0.612 24.45 ± 0.25 ${183}_{-17}^{+14}$
2MASS J2206−2047AB M8+M8 0.0140 0.0160 0.578 ${23.96}_{-0.21}^{+0.23}$ ${188}_{-17}^{+16}$
DENIS J2252−1730AB L4+T3.5 0.0176 0.0182 0.945 ${8.822}_{-0.027}^{+0.026}$ 101 ± 7
Marginal Orbit Determinations
DENIS J1228−1557AB L5.5+L5.5 0.1201 0.0623 0.292 ${50}_{-7}^{+5}$ ${106}_{-19}^{+16}$
2MASS J1847+5522AB M6+M7 0.0346 0.1363 0.216 ${45}_{-3}^{+5}$ ${270}_{-31}^{+26}$
Poorly Constrained Orbits
2MASS J0850+1057AB L6.5+L8.5 0.1167 0.1118 0.310 ${48}_{-6}^{+7}$ 54 ± 8
SDSS J0926+5847AB T3.5+T5 0.3815 0.3761 0.682 ${12.9}_{-1.9}^{+1.3}$ ${38}_{-18}^{+10}$
SDSS J1021−0304AB T0+T5 0.0600 0.1435 0.101 ${86}_{-17}^{+13}$ ${52}_{-7}^{+6}$
SDSS J1534+1615AB T0+T5.5 0.0881 0.4296 0.175 ${58}_{-24}^{+39}$ ${46}_{-7}^{+6}$
2MASS J1750+4424AB M6.5+M8.5 0.1963 0.1523 0.038 ${210}_{-60}^{+40}$ ${190}_{-50}^{+40}$
SDSS J2052−1609AB L8.5+T1.5 0.2291 0.2069 0.195 ${44}_{-14}^{+10}$ ${69}_{-20}^{+14}$

Note. ${\rm{\Delta }}\mathrm{log}{M}_{\mathrm{tot}}$ is defined as the difference between the maximum and minimum values of the 68.3% (1σ) credible interval of the total mass posterior distribution without including the uncertainty in the distance. Likewise, Δe is the difference between the maximum and minimum values of the 1σ credible interval of the eccentricity posterior distribution. Δtobs/P is the time baseline of the resolved astrometry used in the orbit fit divided by the median of the orbital period posterior distribution. For convenience, we also list here the median and 1σ credible intervals of the orbital period and the total mass including the uncertainty on the distance.

Download table as:  ASCIITypeset image

SDSS J0926+5847AB has the worst mass precision ($\delta \mathrm{log}{M}_{\mathrm{tot}}=0.38$ dex) and eccentricity constraint (δe =0.38). Even though it has a good time baseline (${\rm{\Delta }}{t}_{\mathrm{obs}}/P=0.68$), the actual coverage of the orbit on the sky is poor, partly due to the fact that it is seen nearly edge-on ($i=91\buildrel{\circ}\over{.} {7}_{-0.8}^{+0.6}).$ The next worse mass precisions are ${\rm{\Delta }}\mathrm{log}{M}_{\mathrm{tot}}=0.23$ dex for SDSS J2052−1609AB and 0.20 dex for 2MASS J1750+4424AB. The latter has the smallest time baseline of any binary here (${\rm{\Delta }}{t}_{\mathrm{obs}}/P=0.04$), owing to its very long, quite uncertain orbital period (${210}_{-60}^{+40}$ years), although it only has a modestly large δe = 0.15. SDSS J2052−1609AB has a better time baseline (${\rm{\Delta }}{t}_{\mathrm{obs}}/P=0.19$) but a poorly constrained eccentricity (δe = 0.21). 2MASS J0850+1057AB has a relatively poor mass precision (${\rm{\Delta }}\mathrm{log}{M}_{\mathrm{tot}}=0.12$ dex) as well as a poorly constrained eccentricity (δe = 0.11) despite a marginally acceptable time baseline (${\rm{\Delta }}{t}_{\mathrm{obs}}/P=0.31$). We do not consider any of the above orbits sufficiently reliable to be used in the following analysis. SDSS J1534+1615AB and SDSS J1021−0304AB nominally have much better mass precision than any of the above ($\delta \mathrm{log}{M}_{\mathrm{tot}}=0.09$ dex and 0.06 dex, respectively). However, both have poor time baselines with ${\rm{\Delta }}{t}_{\mathrm{obs}}/P=0.17$ and 0.10, respectively, and poorly constrained eccentricities with δe = 0.43 and 0.14, respectively. (For reference, our input prior on eccentricity alone would correspond to δe = 0.68.) We therefore also chose not to use these orbits.

DENIS J1228−1557AB and 2MASS J1847+5522AB are both marginal cases. DENIS J1228−1557AB has a modestly large mass error ($\delta \mathrm{log}{M}_{\mathrm{tot}}=0.12$ dex) but a better constrained eccentricity (δe = 0.06) and longer observational baseline (${\rm{\Delta }}{t}_{\mathrm{obs}}/P=0.29$). On the other hand, 2MASS J1847+5522AB has good mass precision ($\delta \mathrm{log}{M}_{\mathrm{tot}}=0.035$ dex) but a poorly constrained eccentricity (δe = 0.14) and an observational baseline of only ${\rm{\Delta }}{t}_{\mathrm{obs}}/P=0.22$. We conservatively choose to place our cutoffs in orbit quality metrics to exclude these two marginal cases. The next worse mass precision in our sample is LP 415-20AB ($\delta \mathrm{log}{M}_{\mathrm{tot}}=0.06$ dex), but it appears to have a reliably determined orbit, with δe = 0.023 and most of the orbital period covered by observations (${\rm{\Delta }}{t}_{\mathrm{obs}}/P=0.67$). The worst eccentricity constraint for any of the remaining binaries is for 2MASS J1728+3948AB (δe = 0.028), but it has a much better mass precision ($\delta \mathrm{log}{M}_{\mathrm{tot}}=0.0089$ dex) and better orbital coverage (${\rm{\Delta }}{t}_{\mathrm{obs}}/P=0.34$) compared to our excluded orbits. LSPM J1735+2634AB has the worst orbital coverage of our remaining sample (${\rm{\Delta }}{t}_{\mathrm{obs}}/P=0.29$), but its orbit is very well determined ($\delta \mathrm{log}{M}_{\mathrm{tot}}=0.0017$ dex, δe = 0.0075), thanks to our observations serendipitously bracketing its periastron passage.

In summary, we do not use the orbit determinations for 2MASS J0850+1057AB, SDSS J0926+5847AB, DENIS J1228−1557AB, 2MASS J1750+4424AB, 2MASS J1847+5522AB, and SDSS J2052−1609AB in our following analysis. The basis for their exclusion is poor mass precision, poor observational coverage of the orbit, and/or poorly constrained eccentricity that could make our results overly dependent on our uniform eccentricity prior. The remaining sample comprises 23 binaries with orbit quality metrics ranging from $\delta \mathrm{log}{M}_{\mathrm{tot}}\,=0.0017\mbox{--}0.06$ dex, δe = 0.0007–0.028, and ${\rm{\Delta }}{t}_{\mathrm{obs}}/P\,=0.29\mbox{--}5.01.$

4.2. Comparison to Published Orbits

Some of the binaries in our sample have published orbital parameter determinations, either from work of our own or others. By far, the largest overlap of our 31 binary sample is with 14 binaries from Konopacky et al. (2010), followed closely by our own work (13 binaries). In most of these overlapping cases, our new orbit determinations agree with and improve upon previously published work as expected given our longer time baseline and more numerous epochs. For example, compared to Konopacky et al. (2010), our orbital period uncertainties are much smaller, with a median difference in errors of a factor of 7. We now discuss the cases where our new results for astrophysical parameters (i.e., mass, period, and eccentricity, but not viewing angles) differ significantly (≳2σ) from published work.

Gl 569Bab has had a number of published orbit determinations, and despite being a very well-studied system, the reported semimajor axes of the orbit have varied significantly compared to the reported errors. As discussed in Section 3.1 of Dupuy et al. (2010), this is likely mostly due to astrometric calibration issues in data sets that used an early generation of cameras behind the Keck AO system (KCAM and SCAM; Lane et al. 2001), which we did not use here or in our previous work. Our analysis here uses some archival NIRC2 imaging not included in our previous work, which improves the coverage of this short period orbit (P = 2.3707 ± 0.0005 years), and results in a semimajor axis of 93.64 ± 0.14 mas. This sits in between values derived in our prior work (${95.6}_{-1.0}^{+1.1}$ mas Dupuy et al. 2010) and values published by others that largely relied on KCAM and SCAM data: 91.8 ± 1.0 mas (Zapatero Osorio et al. 2004), 90.4 ± 0.7 mas (Simon et al. 2006), and 90.8 ± 0.8 mas (Konopacky et al. 2010). In spite of these different semimajor axes, the dynamical total mass (138 ± 7 ${M}_{\mathrm{Jup}}$ here) varies by <1σ because the error is dominated by the Hipparcos parallax uncertainty (σπ/π = 0.016), not our semimajor axis measurement (σa/a = 0.0016). The improved precision of our new orbit fit is due to more measurements (19 from astrometrically well-calibrated Keck/NIRC2 or HST imaging) that now span 11.87 years (i.e., 5.006 orbital periods).

HD 130948BC has a very well-determined orbit, so small differences in orbital parameters are more statistically significant. In our most recent previous analysis we found $e=0.176\pm 0.006$ (Dupuy & Liu 2011) compared to 0.1627 ± 0.0017 here, a 2.1σ difference. This difference is not astrophysically significant, and perhaps it is simply a statistical fluctuation due to the significant improvement in other orbital parameters (e.g., both a and P are improved by nearly a factor of 20 compared to our original orbit in Dupuy et al. 2009b). Both our current and past values are consistent at <2σ with the eccentricity of 0.16 ± 0.01 from Konopacky et al. (2010).

Our orbit for LP 415-20AB agrees with our previous work but disagrees somewhat with the analysis of Konopacky et al. (2010). This is discussed in detail in Section 2 of Dupuy & Liu (2011) as likely being jointly due to the small number of degrees of freedom in the Konopacky et al. (2010) analysis (1 dof) and one of their measurements being a significant outlier with respect to the rest of the available data. Compared to their orbital parameters, our period of 14.82 ± 0.24 years is longer (compared to 11.5 ± 1.2 years), our semimajor axis of ${96.5}_{-1.4}^{+1.1}$ mas is smaller (compared to 108 ± 24 mas), and our eccentricity of ${0.706}_{-0.012}^{+0.011}$ is smaller (compared to 0.9 ± 0.1). Our inferred total dynamical mass is quite discrepant with that of Konopacky et al. (2010), given that their fitted distance of 21 ± 5 pc (derived by combining astrometry and radial velocities) is 2.7σ different from our parallactic distance of 38.6 ± 1.1 pc. This is likely due both to our different orbit determinations and the fact that their radial velocity measurement (ΔRV = −0.7 ± 1.4 km s−1) disagrees with the prediction from our relative orbit (ΔRV = −2.96 ± 0.16 km s−1). With 12 measurements spanning 9.97 years, our new orbit fit has many more degrees of freedom (17 dof) for the relative orbit fit than previous work, making the resulting orbit parameters more robust and more precise by a factor of 5–17.

Our orbit for 2MASS J1534−2952AB agrees well with Konopacky et al. (2010) but disagrees with Liu et al. (2008). The cause of the discrepancy with Liu et al. (2008) is the choice of eccentricity prior. In that work we adopted a prior of p(e) = e, which deweights smaller eccentricities and rules out circular orbits entirely, whereas we adopt here a more conservative uniform prior p(e) = 1. Our eccentricity posterior here piles up at zero with a 2σ interval of e = 0.000–0.014, and a and P are both correlated with e in the sense that a larger e corresponds to smaller a and P. According to Figure 6 of Liu et al. (2008), that posterior distribution of P as e approaches zero would agree with our orbital period of P = 20.29 ± 0.07 years. According to Figure 8 of Liu et al. (2008), our values of P and a = 213.7 ± 0.5 mas agree well with that Pa posterior distribution in spite of the difference in eccentricity priors. However, both Liu et al. (2008) and Konopacky et al. (2010) used the parallax of 73.6 ± 1.2 mas from Tinney et al. (2003) to convert their angular semimajor axes into physical units and thereby compute dynamical masses. As we discuss in Appendix B.17, our parallax of 63.0 ± 1.1 mas leads to very different dynamical masses than in previous work.

2MASS J2206−2047AB also has an eccentricity posterior that piles up at zero, which differs from the results of our previous work on this system. In Dupuy et al. (2009a) we used a prior of p(e) = e that suppressed low eccentricity orbit solutions and resulted in e = 0.25 ± 0.08. Our new posterior has a 2σ interval of e = 0.000–0.027. As a result of this difference, Dupuy et al. (2009a) found very different values for semimajor axis and period ($a={213}_{-18}^{+24}$ mas, $P={35}_{-5}^{+6}$ years) compared to our new analysis (a = 167.7 ± 0.5 mas, $P={23.96}_{-0.21}^{+0.23}$ years). However, thanks to the strong correlation between a and P, our new values follow the posterior distribution of Dupuy et al. (2009a) as shown in their Figure 6. Therefore, the resulting total dynamical mass was 184 ± 4 ${M}_{\mathrm{Jup}}$ from Dupuy et al. (2009a) but is ${188.3}_{-3.1}^{+2.9}$ ${M}_{\mathrm{Jup}}$ here (assuming a fixed distance of 28.0 pc for the purpose of this comparison). Therefore, the choice of eccentricity prior does not change the best-fit total mass, although it would have affected the uncertainty in the mass, where the uniform e prior used here is more conservative.

SDSS J2052−1609AB has a preliminary orbit determination from Bardalez Gagliuffi et al. (2015) based on combining their data with published astrometry from Stumpf et al. (2011) and three epochs of our data retrieved from the NIRC2 archive. Their posterior eccentricity distribution ($e={0.014}_{-0.010}^{+0.023}$) is significantly different from what we find in our analysis ($e={0.20}_{-0.11}^{+0.09}$, with a 2σ interval of 0.08–0.50). Their quoted uncertainties in other parameters are also generally much smaller than ours, e.g., they find ${M}_{\mathrm{tot}}={86.2}_{-1.8}^{+3.9}$ ${M}_{\mathrm{Jup}}$ while we find ${M}_{\mathrm{tot}}={69}_{-20}^{+14}$ ${M}_{\mathrm{Jup}}$, despite our analysis using more data over a longer time baseline. We also performed a cursory analysis of their astrometry using our MCMC fitter and were unable to reproduce their posterior distributions. Therefore, we speculate that the difference in our results is most likely due to differing MCMC analysis methods. They report that acceptance fractions for their single Metropolis-Hastings sampled chain were typically 0.5%–1%, whereas our analysis using the affine-invariant sampler of emcee had much higher acceptance fractions of 8.7% (Table 6). Their MCMC also included distance as an eighth parameter in addition to the seven orbit parameters that are constrained by the relative astrometry. Since distance is not constrained by the relative astrometry, it appears that its use as a parameter was intended to marginalize over the uncertainty in the Dupuy & Liu (2012) parallax. It is therefore curious that the posterior distribution on distance from Bardalez Gagliuffi et al. (2015) had smaller errors (${30.7}_{-0.4}^{+0.2}$ pc) than the input parallactic distance of 29.5 ± 0.7 pc. Regardless, it is not clear how this would explain the small uncertainties on the orbital parameters or why their MCMC analysis apparently avoided the part of parameter space preferred by our emcee analysis. Our new orbit fit is based on eight epochs, rather than six epochs, spanning 8.58 years instead of 4.55 years, in addition to the fact that our joint Keck+CFHT analysis properly marginalizes over the uncertainty in orbital photocenter motion when determining the parallax. Therefore, we conservatively choose to use our own orbit-fitting results in the following analysis.

Kelu-1AB has an unrefereed orbit determination by Stumpf et al. (2008) based on nine epochs of astrometry spanning 2.84 years. Our orbit based on 13 epochs spanning 16.66 years agrees at ≲1σ with their eccentricity (e = 0.82 ± 0.10) but not their semimajor axis ($a={339}_{-66}^{+129}$ mas) or orbital period ($P={38}_{-6}^{+8}$ years). We find $a={227.9}_{-1.1}^{+0.9}$ mas and P = 24.98 ±0.19 years. Assuming a fixed distance of 20.8 pc for both orbits, we find a total mass of ${M}_{\mathrm{tot}}=180.1\pm 1.1$ ${M}_{\mathrm{Jup}}$ compared to their ${M}_{\mathrm{tot}}={244}_{-76}^{+156}$ ${M}_{\mathrm{Jup}}$, which agrees within 1σ because of their large uncertainties.

4.3. Comparison to Published Parallaxes

In Table 8 we provide a comparison of the parallaxes we measure here with other values in the literature, including our own past works (Dupuy & Liu 2012; Dupuy et al. 2015b). Our previously published parallaxes are not statistically independent of the values given here, but they were derived from somewhat different analysis methods. Namely, in Dupuy & Liu (2012) we did not marginalize over the uncertainty in the orbital parameters or mass ratio in the same way as we have done here, and we used a slightly different method for computing the correction from relative to absolute parallax. It is therefore not surprising that one of the largest discrepancies is for 2MASS J0920+3517AB (2.0σ different from our published parallax), because we previously assumed a model-based mass ratio estimated from the measured flux ratio, but as we discuss in Section 7 and Appendix B.7 this system is likely a hierarchical triple where the fainter component is actually much more massive than the brighter component. The only other ≳2σ difference is for 2MASS J1728+3948AB, where our new data spans 9.1 years (instead of 3.3 years) and with somewhat better parallax phase coverage resulting in a parallax that is 6% (2.5σ) smaller. We suggest this is likely a natural statistical variation given the new better constraints on proper motion in both R.A. and decl. (the parallax amplitude in decl. is almost as large as R.A. for this object).

Table 8.  Comparison of Parallaxes and Proper Motions to Literature Values

Object Absolute Parallax (mas) ΔProper Motion (mas/years)a References
  Here Lit. Difference R.A. Decl.  
LP 349-25AB 69.2 ± 0.9 75.8 ± 1.6 1.9σ 10% 7 18 Gate09
SDSS J0423−0414AB 71.1 ± 0.8 65.9 ± 1.7 −1.5σ −7% −1.3 ± 2.9 −6.2 ± 1.9 Vrba04
SDSS J0423−0414AB 71.1 ± 0.8 67.5 ± 2.3 −0.6σ −5% 0.6 ± 1.4 −6.0 ± 1.3 Wein16
2MASS J0700+3157AB 88.6 ± 0.9 82.0 ± 2.0 −1.4σ −7% 7.1 ± 1.0 7.2 ± 1.1 Thor03
LHS 1901AB 76.4 ± 1.1 78 ± 3 0.1σ 2% 2 −8 Lepi09
2MASS J0746+2000AB 80.9 ± 0.8 86 ± 5 0.2σ 6% 9 ± 5 −11 ± 5 Fahe12
2MASS J0746+2000AB 80.9 ± 0.8 81.9 ± 0.3 1.3σ 1% −8.9 ± 0.9 −4.8 ± 1.2 Dahn02
2MASS J0850+1057AB 31.4 ± 0.6 39 ± 4 0.6σ 24% 3 ± 3 7.8 ± 2.7 Dahn02
2MASS J0850+1057AB 31.4 ± 0.6 26 ± 4 −0.3σ −17% 1 ± 7 −5 ± 3 Vrba04
2MASS J0850+1057AB 31.4 ± 0.6 35 ± 8 0.1σ 11% 3 ± 6 −23 ± 6 Fahe11
SDSS J1021−0304AB 33.7 ± 1.2 41 ± 11 0.1σ 21% 23 ± 16 −1 ± 9 Vrba04
SDSS J1021−0304AB 33.7 ± 1.2 34 ± 5 0.0σ 2% 16 ± 9 5 ± 4 Tinn03
LHS 2397aAB 68.2 ± 1.1 65.8 ± 2.0 −0.5σ −4% −10.8 ± 1.7 12.0 ± 1.6 Diet14
LHS 2397aAB 68.2 ± 1.1 70.0 ± 2.1 0.3σ 3% 28.9 ± 1.8 8.5 ± 1.5 Mone92
DENIS J1228−1557AB 48.0 ± 1.7 49.4 ± 1.9 0.2σ 3% 30 ± 30 31 ± 27 Dahn02
Kelu-1AB 48.0 ± 2.2 53.6 ± 2.0 0.6σ 12% 15 ± 8 13 ± 5 Dahn02
Kelu-1AB 48.0 ± 2.2 51.8 ± 1.2 0.6σ 8% 1 ± 8 −7 ± 5 Wein16
2MASS J1404−3159AB 42.5 ± 1.1 49 ± 3 0.5σ 16% 3.7 ± 2.0 5.8 ± 2.5 Maro13
2MASS J1404−3159AB 42.5 ± 1.1 40 ± 6 −0.1σ −6% 1 ± 5 7 ± 5 Fahe12
2MASS J1534−2952AB 63.0 ± 1.1 73.6 ± 1.2 3.9σ 17% 4.7 ± 1.1 11.1 ± 2.4 Tinn03
2MASS J1728+3948AB 36.4 ± 0.6 42 ± 3 0.4σ 14% 2 ± 6 −4 ± 5 Vrba04
SDSS J2052−1609AB 33.7 ± 0.7 27 ± 9 −0.1σ −21% 10 ± 11 21 ± 10 Fahe12
2MASS J2132+1341AB 35.6 ± 0.6 30 ± 8 −0.1σ −16% 0 ± 12 −15 ± 11 Fahe12
2MASS J2206−2047AB 35.8 ± 1.0 38 ± 3 0.1σ 5% −2 ± 7 5 ± 5 Cost06
Previously Published CFHT Parallaxes
LP 349-25AB 69.2 ± 0.9 69.6 ± 0.9 0.2σ 1% −3.9 ± 2.0 5.0 ± 2.0 Dupu12
SDSS J0423−0414AB 71.1 ± 0.8 72.1 ± 1.1 0.5σ 1% −6.0 ± 1.1 3.1 ± 1.3 Dupu12
2MASS J0700+3157AB 88.6 ± 0.9 86.7 ± 1.2 −0.8σ −2% 20.0 ± 1.0 −1.6 ± 1.1 Dupu12
LHS 1901AB 76.4 ± 1.1 74.2 ± 1.0 −1.0σ −3% −1.8 ± 1.5 0 ± 3 Dupu12
2MASS J0746+2000AB 80.9 ± 0.8 81.1 ± 0.9 0.1σ 0% −0.8 ± 1.1 0.4 ± 1.1 Dupu12
2MASS J0850+1057AB 31.4 ± 0.6 30.1 ± 0.8 −1.2σ −4% 2.9 ± 2.4 2.8 ± 1.5 Dupu12
2MASS J0920+3517AB 32.3 ± 0.6 34.4 ± 0.8 2.0σ 6% −2.3 ± 1.5 2.6 ± 1.3 Dupu12
SDSS J0926+5847AB 43.4 ± 1.0 43.7 ± 1.1 0.1σ 1% −1.9 ± 1.4 3.3 ± 1.3 Dupu12
2MASS J1017+1308AB 32.3 ± 1.1 30.2 ± 1.4 −0.6σ −6% 0.9 ± 2.0 5.0 ± 1.9 Dupu12
SDSS J1021−0304AB 33.7 ± 1.2 29.9 ± 1.3 −1.2σ −11% 24 ± 8 −3.2 ± 2.1 Dupu12
SDSS J1052+4422AB 38.1 ± 0.6 38.4 ± 0.7 0.3σ 1% −2 ± 3 −2 ± 3 Dupu15
LHS 2397aAB 68.2 ± 1.1 73.0 ± 2.1 0.9σ 7% 7.0 ± 3.0 −2.6 ± 2.1 Dupu12
DENIS J1228−1557AB 48.0 ± 1.7 44.8 ± 1.8 −0.5σ −7% 30 ± 30 25 ± 27 Dupu12
Kelu-1AB 48.0 ± 2.2 49.7 ± 2.4 0.2σ 4% 1 ± 9 −2 ± 5 Dupu12
2MASS J1404−3159AB 42.5 ± 1.1 42.1 ± 1.1 −0.2σ −1% 10.9 ± 1.1 11.4 ± 1.6 Dupu12
SDSS J1534+1615AB 27.6 ± 0.9 24.9 ± 1.1 −1.4σ −10% −1.1 ± 1.7 4.2 ± 1.2 Dupu12
2MASS J1534−2952AB 63.0 ± 1.1 62.4 ± 1.3 −0.2σ −1% 2.2 ± 1.2 2.2 ± 2.0 Dupu12
2MASS J1728+3948AB 36.4 ± 0.6 38.7 ± 0.7 2.5σ 6% 1.0 ± 0.8 3.6 ± 1.0 Dupu12
LSPM J1735+2634AB 64.7 ± 0.8 66.7 ± 1.4 0.7σ 3% −1.3 ± 1.1 −3.9 ± 1.5 Dupu12
2MASS J1750+4424AB 32.4 ± 0.9 30.3 ± 1.0 −1.1σ −6% 30 ± 60 −7 ± 29 Dupu12
2MASS J1847+5522AB 27.4 ± 0.9 29.8 ± 1.1 1.2σ 9% −0.9 ± 1.1 2.3 ± 1.6 Dupu12
SDSS J2052−1609AB 33.7 ± 0.7 33.9 ± 0.8 0.1σ 1% −0.7 ± 0.8 2.9 ± 1.5 Dupu12
2MASS J2132+1341AB 35.6 ± 0.6 36.0 ± 0.7 0.4σ 1% 3.0 ± 1.4 3.3 ± 0.9 Dupu12
2MASS J2140+1625AB 30.4 ± 0.9 32.5 ± 1.1 1.1σ 7% 0.9 ± 1.8 −2.5 ± 1.7 Dupu12
2MASS J2206−2047AB 35.8 ± 1.0 35.7 ± 1.2 −0.0σ 0% −7.7 ± 1.9 3.9 ± 1.8 Dupu12
DENIS J2252−1730AB 63.1 ± 1.4 63.2 ± 1.6 0.0σ 0% −9.2 ± 2.0 6 ± 4 Dupu12

Note.

aThe difference between our absolute proper motions and literature values that are typically reported as relative.

References. Cost06—Costa et al. (2006), Dahn02—Dahn et al. (2002), Diet14—Dieterich et al. (2014), Dupu12—Dupuy & Liu (2012), Dupu15—Dupuy et al. (2015b), Fahe11—Faherty et al. (2011), Fahe12—Faherty et al. (2012), Gate09—Gatewood & Coban (2009), Lepi09—Lépine et al. (2009), Maro13—Marocco et al. (2013), Mone92—Monet et al. (1992), Thor03—Thorstensen & Kirkpatrick (2003), Tinn03—Tinney et al. (2003), Vrba04—Vrba et al. (2004), Wein16—Weinberger et al. (2016).

Download table as:  ASCIITypeset image

Among other literature parallaxes, the largest discrepancy is for 2MASS J1534−2952AB (3.9σ). As we have previously discussed in Dupuy & Liu (2012), this is likely due to the uncertainty in the Tinney et al. (2003) value of 73.6 ± 1.2 mas being somewhat underestimated. Our value of 63.0 ± 1.1 mas significantly changes the inferred dynamical mass since ${M}_{\mathrm{tot}}\propto {d}^{3}$, resulting in a factor of 1.6 increase in mass. The next largest discrepancies are 1.3σ−1.9σ for LP 349-25AB, SDSS J0423−0414AB, 2MASS J0700+3157AB, and 2MASS J0746+2000AB and the remaining 19 comparisons agree within ≤0.6σ. Two binaries have parallaxes in the literature more precise than our CFHT values. Kelu-1AB has a parallax of 53.6 ± 2.0 mas from Dahn et al. (2002) and 51.75 ± 1.16 mas from Weinberger et al. (2016) that are both more precise but likely less conservative than our value of 49.8 ± 2.2 mas because we marginalize over photocenter orbital motion. All of these values agree with each other at 0.8σ–1.3σ, but using the different parallaxes results in quite different dynamical masses, and we therefore exclude Kelu-1AB from our further mass analysis (see Appendix B.13 for a discussion). 2MASS J0746+2000AB has a remarkably precise parallax of 81.9 ± 0.3 mas from the USNO optical program (Dahn et al. 2002). An updated USNO analysis that accounts for the photocenter orbital motion of 2MASS J0746+2000AB (Harris et al. 2015) results in a parallax of 81.24 ± 0.25 mas (H. Harris 2015, private communication). Given the longer time baseline and higher precision of the USNO parallax, we use their value in our following analysis.

5. Empirically Determined Properties

5.1. Spectral Types and Magnitudes

Most binaries in our sample have spectral types derived through spectral decomposition from Dupuy & Liu (2012), using integrated-light NIR spectra and resolved NIR photometry. For most of these we simply adopt the same spectral types as in that work, but a few binaries were not in that sample or have updated resolved photometry that warrants new analysis. We use the same method described in Section 5.2 of Dupuy & Liu (2012). Briefly, we pair spectra from a template library and find the optimal scaling ratio for each pair that best matches the observed integrated-light spectrum. We compute synthetic photometry for these pairs and then cull all pairings that do not agree with the observed resolved photometry, p(χ2) < 0.05. We then examine the remaining pairs, ranked by how well they match the observed spectrum, and assign types and errors on types that best represent the results. A summary of all available magnitudes for our sample binaries, both integrated-light and resolved, is given in Table 9. We report magnitudes on both the 2MASS and MKO photometric systems if data exist in either system by using the empirical relations between photometric system conversion and absolute magnitude that we derived in Appendix A.1.

Table 9.  Apparent Magnitudes for the Sample

Filter Integrated ${\rm{\Delta }}m={m}_{{\rm{B}}}-{m}_{{\rm{A}}}$ mA mB References
LP 349-25AB
MKO J 10.563 ± 0.022 0.350 ± 0.030 11.155 ± 0.025 11.505 ± 0.028 Cutr03, Dupu10, Dupu12
MKO H 10.005 ± 0.023 0.326 ± 0.011 10.607 ± 0.023 10.933 ± 0.024 Cutr03, Dupu10, Dupu12
MKO K 9.540 ± 0.017 0.307 ± 0.008 10.150 ± 0.017 10.457 ± 0.018 Cutr03, Dupu10, Dupu12
2MASS J 10.615 ± 0.025 0.346 ± 0.037 11.208 ± 0.029 11.555 ± 0.032 Cutr03, Dupu10, Dupu12, Dupu17
2MASS H 9.976 ± 0.023 0.320 ± 0.012 10.580 ± 0.024 10.900 ± 0.024 Cutr03, Dupu10, Dupu12, Dupu17
2MASS KS 9.569 ± 0.017 0.318 ± 0.007 10.174 ± 0.017 10.492 ± 0.017 Cutr03, Dupu10

References. Burg03b—Burgasser et al. (2003b), Burg06b—Burgasser et al. (2006b), Burg11—Burgasser et al. (2011), Chiu06—Chiu et al. (2006), Clos03—Close et al. (2003), Cutr03—Cutri et al. (2003), Dupu09a—Dupuy et al. (2009a), Dupu09b—Dupuy et al. (2009b), Dupu09c—Dupuy et al. (2009c), Dupu10—Dupuy et al. (2010), Dupu12—Dupuy & Liu (2012), Dupu14—Dupuy et al. (2014), Dupu15—Dupuy et al. (2015b), Dupu17—this work, Gizi03—Gizis et al. (2003), Knap04—Knapp et al. (2004), Kono10—Konopacky et al. (2010), Lane01—Lane et al. (2001), Legg00—Leggett et al. (2000), Legg02a—Leggett et al. (2002), Liu08—Liu et al. (2008), Reid06a—Reid et al. (2006a), Stum11—Stumpf et al. (2011).

Only a portion of this table is shown here to demonstrate its form and content. A machine-readable version of the full table is available.

Download table as:  DataTypeset image

LP 415-20AB and 2MASS J1047+4026AB were not included in the Dupuy & Liu (2012) sample. LP 415-20AB has an integrated-light optical spectral type of M7.5 (Gizis et al. 2000a), but we find that primary templates as early as M5 and as late as M7 provide good matches to the infrared SpeX SXD spectrum and resolved photometry, while the best secondary templates range from M7.5–M8.5. Therefore, we adopt types of M6.0 ± 1.0 for LP 415-20A and M8.0 ± 0.5 for LP 415-20B. 2MASS J1047+4026AB has an integrated-light optical type of M8 (Gizis et al. 2000b), and we find types of M8.0 ± 0.5 and L0.0 ± 1.0 in our spectral decomposition. In this work we have added new flux ratios in the CH4S and K bands for 2MASS J1404−3159AB that allow us to refine our analysis, and we find the same spectral types as we did in Dupuy & Liu (2012), L9 ± 1 and T5.0 ± 0.5. We have added J-, H-, and K-band flux ratios for 2MASS J2140+1625AB, resulting in the same primary type (M8.0 ± 0.5) but an updated secondary type of L0.5 ± 1.0 (was M9.5 ± 0.5). Finally, for DENIS J2252−1730AB we have new flux ratios in the Y, J, H, CH4S, and K bands. This allows us to improve the spectral types to L4.0 ± 1.0 (was L4.5 ± 1.5) and T3.5 ± 0.5 (was T3.5 ± 1.0).

5.2. Individual Masses

Our joint analysis of resolved relative astrometry and unresolved absolute astrometry provides direct measurements of the relative orbit, parallax, and photocenter orbit. The first two of these give the total system mass directly. When the flux ratio of the binary is known, then the photocenter orbit can also be used to directly measure the individual component masses. Our absolute astrometry from CFHT/WIRCam is all in either the J or ${K}_{{\rm{H}}2}$ band, and we have resolved J- and K-band photometry for all of our binaries. Therefore, we can derive individual masses for any binary with a sufficiently well-constrained photocenter orbit semimajor axis (${a}_{\mathrm{phot}}$). As noted above, we assume a uniform prior in the ratio of the photocenter semimajor axis to the true semimajor axis (${a}_{\mathrm{phot}}$/a). If we define the ratio of the secondary-to-total mass as $\mu \equiv {M}_{2}/({M}_{1}+{M}_{2})$ and the ratio of secondary to total flux as $\beta \equiv {F}_{2}/({F}_{1}+{F}_{2})$, then ${a}_{\mathrm{phot}}/a=\mu -\beta $. An unbounded and uniform prior on ${a}_{\mathrm{phot}}$/a and β, as we have assumed, can in principle allow for unphysical values of μ < 0. This is simply a consequence of not using any information about the flux ratio in our astrometric analysis, motivated by the fact that improved flux ratios could easily be obtained in the future. Therefore, not all astrometric solutions result in meaningful constraints on the individual mass ratios, but in our sample of well-determined orbits only Kelu-1AB has an ${a}_{\mathrm{phot}}$ uncertainty sufficiently large to encompass a wide range of unphysical values.

In order to derive flux ratios, we used the values of ΔJ and ΔK reported in Table 9. We use the MKO J-band data directly, while for the narrow-band ${K}_{{\rm{H}}2}$ filter on WIRCam we must derive a correction to be applied to our broadband K measurements. As described in Appendix A.2, we derived third-order polynomial relations between the K-band absolute magnitude and both ${K}_{\mathrm{MKO}}-{K}_{{\rm{H}}2}$ color and ${K}_{2\mathrm{MASS}}-{K}_{{\rm{H}}2}$ color. These relations have an rms scatter of 0.021 mag and 0.025 mag, respectively. When applying the offset from these relations we added these values of the rms scatter in quadrature to the flux ratio uncertainty in any case where our observed ΔK was more than 2× larger than the scatter. In other words, for binaries with flux ratios within ≲0.05 mag of unity, we did not add this uncertainty in quadrature.

Table 10 summarizes all of the photocenter orbit sizes and flux ratios measured in our analysis, including systems for which the dynamical masses are not reliable. Table 11 gives the resulting individual masses we derive from our dynamical total masses and photocenter mass ratios (19 systems) or from other information (three systems). Two of the three binaries in our relative orbit sample that do not have absolute astrometry from CFHT have stellar companions with Hipparcos distances (Gl 417BC and HD 130948BC). We provide the model-derived individual masses in Table 11 for reference, but we do not consider them as part of the following analysis of individual masses. The third binary with an orbit but no CFHT data is Gl 569Bab, and for this system we use the mass ratio of M1/M2 = 1.4 ± 0.3 derived from a joint analysis of all available radial velocity data by Konopacky et al. (2010, see their Section 5.4.6). Gl 569Bab and LHS 1070BC (Köhler et al. 2012) are the only two ultracool binary systems that previously had individual mass determinations in the literature not from our own work. Our sample of 38 objects with individual masses therefore increases the sample size by an order of magnitude.

Table 10.  Summary of Key Parameters from Orbit Analysis

System ${M}_{\mathrm{tot}}$ q e P a ${a}_{\mathrm{phot}}$/a β
  (${M}_{\mathrm{Jup}}$) M2/M1   (days) (au)   $\equiv {F}_{2}/({F}_{1}+{F}_{2})$ a
LP 349-25AB ${166}_{-7}^{+6}$ ${0.941}_{-0.030}^{+0.029}$ ${0.0468}_{-0.0018}^{+0.0019}$ 2812 ± 5 2.109 ± 0.027 0.056 ± 0.008 0.4286 ± 0.0018
LP 415-20AB ${248}_{-29}^{+26}$ ${0.59}_{-0.12}^{+0.10}$ ${0.706}_{-0.012}^{+0.011}$ 5410 ± 90 3.73 ± 0.12 0.00 ± 0.03 0.370 ± 0.028
SDSS J0423−0414AB 83 ± 3 0.62 ± 0.04 ${0.272}_{-0.007}^{+0.008}$ ${4491}_{-22}^{+21}$ ${2.291}_{-0.028}^{+0.027}$ −0.023 ± 0.006 0.404 ± 0.013
2MASS J0700+3157AB ${141}_{-5}^{+4}$ 1.08 ± 0.05 ${0.017}_{-0.007}^{+0.005}$ 8720 ± 190 4.25 ± 0.08 0.310 ± 0.011 0.208 ± 0.005
LHS 1901AB ${213}_{-10}^{+9}$ ${0.87}_{-0.11}^{+0.09}$ 0.8304 ± 0.0009 5921 ± 29 3.76 ± 0.06 −0.009 ± 0.028 0.4750 ± 0.0016
2MASS J0746+2000AB ${160.8}_{-1.7}^{+1.8}$ ${0.952}_{-0.027}^{+0.026}$ 0.4854 ± 0.0003 4652 ± 11 2.919 ± 0.009 0.069 ± 0.007 0.4186 ± 0.0018
2MASS J0850+1057AB 54 ± 8b ${0.06}_{-0.06}^{+0.05}$ b ${17600}_{-2300}^{+2700}$ b ${4.98}_{-0.33}^{+0.25}$ b 0.04 ± 0.12 0.225 ± 0.021
2MASS J0920+3517AB 187 ± 11 ${1.63}_{-0.09}^{+0.08}$ ${0.180}_{-0.007}^{+0.006}$ 2651 ± 3 2.11 ± 0.04 0.183 ± 0.010 0.437 ± 0.006
SDSS J0926+5847AB ${38}_{-18}^{+10}$ b ${0.35}_{-0.16}^{+0.22}$ b ${4700}_{-700}^{+500}$ b ${1.80}_{-0.21}^{+0.14}$ b −0.07 ± 0.08 0.452 ± 0.030
2MASS J1017+1308AB ${156}_{-18}^{+14}$ 0.92 ± 0.08 0.158 ± 0.010 6790 ± 80 3.72 ± 0.13 0.026 ± 0.022 0.454 ± 0.004
SDSS J1021−0304AB ${52}_{-7}^{+6}$ b 0.38 ± 0.07b ${31000}_{-6000}^{+5000}$ b ${7.2}_{-0.8}^{+0.7}$ b −0.9 ± 0.3 0.5239 ± 0.0028
2MASS J1047+4026AB ${178}_{-12}^{+11}$ 0.82 ± 0.06 0.7485 ± 0.0013 ${2397}_{-10}^{+11}$ 1.94 ± 0.04 0.019 ± 0.015 0.432 ± 0.011
SDSS J1052+4422AB ${90}_{-5}^{+4}$ 0.78 ± 0.07 ${0.1399}_{-0.0023}^{+0.0022}$ 3144 ± 9 1.86 ± 0.03 −0.165 ± 0.008 0.602 ± 0.020
Gl 417BC ${99.2}_{-3.3}^{+3.0}$ 0.105 ± 0.003 ${5714}_{-30}^{+29}$ 2.851 ± 0.029
LHS 2397aAB ${159}_{-8}^{+7}$ ${0.706}_{-0.028}^{+0.027}$ 0.351 ± 0.003 ${5248}_{-17}^{+18}$ 3.15 ± 0.05 0.344 ± 0.009 0.0693 ± 0.0025
DENIS J1228−1557AB ${106}_{-19}^{+16}$ b ${0.089}_{-0.035}^{+0.027}$ b ${18300}_{-2400}^{+1900}$ b ${6.36}_{-0.35}^{+0.29}$ b 0.9 ± 1.1 0.476 ± 0.008
Kelu-1AB ${180}_{-26}^{+22}$ c 0.709 ± 0.005 9120 ± 70 ${4.75}_{-0.22}^{+0.21}$ −0.7 ± 0.9 0.344 ± 0.010
2MASS J1404−3159AB ${120}_{-13}^{+11}$ 0.84 ± 0.06 0.825 ± 0.005 6030 ± 80 ${3.15}_{-0.11}^{+0.09}$ −0.165 ± 0.016 0.622 ± 0.006
HD 130948BC ${115.4}_{-2.1}^{+2.2}$ 0.1627 ± 0.0017 3656 ± 4 ${2.226}_{-0.013}^{+0.014}$
Gl 569Bab 138 ± 7 0.3186 ± 0.0010 865.89 ± 0.20 0.904 ± 0.015
SDSS J1534+1615AB ${46}_{-7}^{+6}$ b ${0.22}_{-0.22}^{+0.21}$ b ${21000}_{-9000}^{+14000}$ b ${5.3}_{-1.6}^{+2.3}$ b 0.07 ± 0.10 0.5411 ± 0.0011
2MASS J1534−2952AB 99 ± 5 ${0.95}_{-0.16}^{+0.13}$ ${0.0027}_{-0.0027}^{+0.0028}$ ${7434}_{-21}^{+18}$ 3.40 ± 0.06 0.02 ± 0.04 0.463 ± 0.003
2MASS J1728+3948AB ${140}_{-8}^{+7}$ ${0.93}_{-0.13}^{+0.11}$ ${0.015}_{-0.015}^{+0.013}$ ${14900}_{-400}^{+600}$ ${6.09}_{-0.22}^{+0.17}$ 0.04 ± 0.03 0.439 ± 0.012
LSPM J1735+2634AB 187 ± 7 ${0.868}_{-0.025}^{+0.023}$ 0.497 ± 0.004 7910 ± 90 ${4.37}_{-0.06}^{+0.07}$ 0.077 ± 0.007 0.3872 ± 0.0025
2MASS J1750+4424AB ${190}_{-50}^{+40}$ b ${0.73}_{-0.07}^{+0.09}$ b ${78000}_{-22000}^{+15000}$ b ${20}_{-4}^{+3}$ b −0.8 ± 2.1 0.354 ± 0.008
2MASS J1847+5522AB ${270}_{-31}^{+26}$ b ${0.09}_{-0.09}^{+0.05}$ b ${16500}_{-1200}^{+2000}$ b ${8.0}_{-0.6}^{+0.7}$ b 0.05 ± 0.08 0.439 ± 0.005
SDSS J2052−1609AB ${69}_{-20}^{+14}$ b ${0.20}_{-0.11}^{+0.09}$ b ${16000}_{-5000}^{+4000}$ b ${5.0}_{-0.6}^{+0.5}$ b 0.06 ± 0.05 0.490 ± 0.013
2MASS J2132+1341AB ${128}_{-8}^{+7}$ ${0.88}_{-0.05}^{+0.04}$ ${0.315}_{-0.005}^{+0.004}$ 3920 ± 60 2.42 ± 0.04 0.155 ± 0.010 0.314 ± 0.008
2MASS J2140+1625AB ${183}_{-17}^{+14}$ ${0.60}_{-0.08}^{+0.07}$ 0.196 ± 0.007 8930 ± 90 4.71 ± 0.14 0.042 ± 0.025 0.334 ± 0.016
2MASS J2206−2047AB ${188}_{-17}^{+16}$ ${0.84}_{-0.10}^{+0.09}$ 0.015 ± 0.008 8750 ± 80 ${4.69}_{-0.14}^{+0.13}$ −0.033 ± 0.028 0.4885 ± 0.0023
DENIS J2252−1730AB 101 ± 7 ${0.70}_{-0.09}^{+0.08}$ 0.334 ± 0.009 ${3222}_{-10}^{+9}$ 1.95 ± 0.04 0.081 ± 0.027 0.330 ± 0.008

Notes.

aThis is the ratio of the secondary's flux to the total flux in the bandpass used for our CFHT/WIRCam absolute astrometry observations, which was either a broadband J (MKO) or a narrowband filter in the K band centered at 2.122 μm. bWe do not use these orbit determinations in our analysis given their questionable quality (see Section 4.1). cThis is the total mass of Kelu-1AB using our CFHT parallax, but we do not use it in our analysis given that parallaxes from the literature differ from ours at a level sufficient to result in significantly different physical properties.

Download table as:  ASCIITypeset image

Table 11.  Summary of Properties for Dynamical Mass Sample

  Primary Component Secondary Component  
System M (${M}_{\mathrm{Jup}}$) $\mathrm{log}({L}_{\mathrm{bol}}/{L}_{\odot })$ SpT ${T}_{\mathrm{eff}}$ (K) M (${M}_{\mathrm{Jup}}$) $\mathrm{log}({L}_{\mathrm{bol}}/{L}_{\odot })$ SpT ${T}_{\mathrm{eff}}$ (K) Age (Gyr)
LP 349-25AB 85 ± 4 $-{3.075}_{-0.026}^{+0.027}$ M7 ± 1.0 ${2729}_{-27}^{+26}$ 80 ± 3 −3.198 ± 0.027 M8 ± 1.0 ${2629}_{-27}^{+29}$ ${0.271}_{-0.029}^{+0.022}$
LP 415-20AB ${156}_{-18}^{+17}$ −3.010 ± 0.030 M6 ± 1.0 ${92}_{-18}^{+16}$ $-{3.260}_{-0.040}^{+0.030}$ M8 ± 0.5 ${2640}_{-30}^{+40}$
SDSS J0423−0414AB ${51.6}_{-2.5}^{+2.3}$ −4.41 ± 0.04 L6.5 ± 1.5 ${1430}_{-40}^{+30}$ ${31.8}_{-1.6}^{+1.5}$ −4.72 ± 0.07 T2 ± 0.5 1200 ± 40 ${0.81}_{-0.09}^{+0.07}$
2MASS J0700+3157AB 68.0 ± 2.6 −3.95 ± 0.04 L3 ± 1.0 1860 ± 40 ${73.3}_{-3.0}^{+2.9}$ −4.45 ± 0.04 L6.5 ± 1.5 ${0.76}_{-0.14}^{+0.09}$
LHS 1901AB 113 ± 8 $-{3.005}_{-0.027}^{+0.026}$ M7 ± 1.0 2844 ± 16 99 ± 7 $-{3.046}_{-0.027}^{+0.026}$ M7 ± 1.0 ${2820}_{-10}^{+20}$
2MASS J0746+2000AB ${82.4}_{-1.5}^{+1.4}$ $-{3.596}_{-0.025}^{+0.028}$ L0 ± 0.5 ${2318}_{-23}^{+24}$ 78.4 ± 1.4 $-{3.777}_{-0.027}^{+0.028}$ L1.5 ± 0.5 ${2134}_{-25}^{+26}$
2MASS J0920+3517AB 71 ± 5 −4.270 ± 0.030 L5.5 ± 1.0 ${1621}_{-30}^{+32}$ ${116}_{-8}^{+7}$ −4.340 ± 0.030 L9 ± 1.5
2MASS J1017+1308AB ${81}_{-11}^{+10}$ −3.78 ± 0.04 L1.5 ± 1.0 2090 ± 50 75 ± 7 −3.84 ± 0.04 L3 ± 1.0 ${2040}_{-50}^{+60}$
2MASS J1047+4026AB ${97}_{-7}^{+6}$ −3.27 ± 0.04 M8 ± 0.5 ${2640}_{-40}^{+50}$ 80 ± 6 −3.39 ± 0.05 L0 ± 1.0 ${2510}_{-50}^{+60}$
SDSS J1052+4422AB 51 ± 3 −4.51 ± 0.04 L6.5 ± 1.5 ${1366}_{-29}^{+25}$ ${39.4}_{-2.7}^{+2.6}$ −4.64 ± 0.07 T1.5 ± 1.0 1270 ± 40 ${1.04}_{-0.15}^{+0.14}$
Gl 417BCa ${51.5}_{-1.8}^{+1.7}$ −4.132 ± 0.030 L4.5 ± 1.0 ${1639}_{-31}^{+29}$ 47.7 ± 1.9 −4.220 ± 0.030 L6 ± 1.0 ${1560}_{-26}^{+29}$ ${0.49}_{-0.04}^{+0.03}$
LHS 2397aAB 93 ± 4 −3.34 ± 0.04 M8 ± 0.5 2560 ± 50 66 ± 4 −4.48 ± 0.04 1440 ± 40 ${2.8}_{-1.5}^{+2.1}$
2MASS J1404−3159AB 65 ± 6 $-{4.52}_{-0.05}^{+0.06}$ L9 ± 1.0 ${1400}_{-50}^{+40}$ ${55}_{-7}^{+6}$ $-{4.87}_{-0.07}^{+0.08}$ T5 ± 0.5 1190 ± 50 ${3.0}_{-1.3}^{+0.8}$
HD 130948BCa ${59.8}_{-2.1}^{+2.0}$ −3.85 ± 0.06 L4 ± 1.0 ${1920}_{-60}^{+70}$ ${55.6}_{-1.9}^{+2.0}$ −3.96 ± 0.06 L4 ± 1.0 ${1800}_{-70}^{+50}$ 0.44 ± 0.04
Gl 569Babb ${80}_{-8}^{+9}$ −3.440 ± 0.030 M8.5 ± 0.5 2420 ± 40 ${58}_{-9}^{+7}$ −3.670 ± 0.030 M9 ± 0.5 2170 ± 50 ${0.44}_{-0.06}^{+0.05}$
2MASS J1534−2952AB 51 ± 5 −4.91 ± 0.07 T4.5 ± 0.5 ${1150}_{-50}^{+40}$ 48 ± 5 −4.99 ± 0.07 T5 ± 0.5 ${1100}_{-50}^{+40}$ ${3.0}_{-0.5}^{+0.4}$
2MASS J1728+3948AB 73 ± 7 $-{4.29}_{-0.05}^{+0.04}$ L5 ± 1.0 1600 ± 40 67 ± 5 −4.49 ± 0.04 L7 ± 1.0 1440 ± 40 ${3.4}_{-2.1}^{+2.8}$
LSPM J1735+2634AB 100 ± 4 −3.221 ± 0.028 M7.5 ± 0.5 ${2677}_{-26}^{+28}$ 87 ± 3 $-{3.445}_{-0.028}^{+0.029}$ L0 ± 1.0 2462 ± 30
2MASS J2132+1341AB 68 ± 4 −4.22 ± 0.05 L4.5 ± 1.5 ${1660}_{-40}^{+50}$ 60 ± 4 $-{4.50}_{-0.04}^{+0.05}$ L8.5 ± 1.5 ${1400}_{-40}^{+30}$ ${1.44}_{-0.37}^{+0.26}$
2MASS J2140+1625AB ${114}_{-12}^{+10}$ −3.20 ± 0.04 M8 ± 0.5 2680 ± 40 ${69}_{-9}^{+8}$ −3.51 ± 0.04 L0.5 ± 1.0 ${2410}_{-50}^{+60}$
2MASS J2206−2047AB ${102}_{-11}^{+10}$ $-{3.220}_{-0.040}^{+0.030}$ M8 ± 0.5 2670 ± 40 ${86}_{-10}^{+8}$ $-{3.260}_{-0.030}^{+0.040}$ M8 ± 0.5 ${2650}_{-30}^{+40}$
DENIS J2252−1730AB 59 ± 5 $-{4.26}_{-0.04}^{+0.05}$ L4 ± 1.0 1590 ± 50 41 ± 4 $-{4.76}_{-0.07}^{+0.08}$ T3.5 ± 0.5 ${1210}_{-40}^{+50}$ ${1.11}_{-0.22}^{+0.19}$

Notes. The effective temperatures quoted here are derived from the SM08 models for objects cooler than 2100 K and from the Baraffe et al. (2015, hereafter BHAC) models for objects hotter than 2100 K. System ages are only reported in cases where the observations provide a meaningful constraint, and most of the quoted values are from our total-mass analysis using the SM08 models. The exceptions are LP 349-25AB and Gl 569Bab (total-mass analysis using the BHAC models), 2MASS J0700+3157AB (individual-mass analysis of the primary using SM08 models), and LHS 2397aAB (individual-mass analysis of the secondary using SM08 models). The individual masses given here are directly measured in this work unless otherwise noted.

aThe individual masses quoted for Gl 417BC and HD 130948BC are not measured directly but rather are inferred from evolutionary models in our total-mass analysis. The component effective temperatures are also from our total-mass analysis. bThe individual masses quoted for Gl 569Bab are computed from our total mass and the mass ratio from Konopacky et al. (2010). The component effective temperatures are from our total-mass analysis.

Download table as:  ASCIITypeset image

Figure 3 shows all of our individual masses, as well as those from the literature, as a function of spectral type. In the analysis that follows, we will use this information to discuss the substellar boundary. Here we simply note that in most cases, our individual masses do not grossly deviate from rough astrophysical expectations, e.g., nearly equal flux systems have mass ratios near unity and vice versa. There are a few remarkable exceptions, such as the secondary 2MASS J0920+3517B (L9.0 ± 1.5) that has a mass of ${116}_{-8}^{+7}$ ${M}_{\mathrm{Jup}}$, well in excess of the expected hydrogen fusion limit (≈70 ${M}_{\mathrm{Jup}}$; Section 7.1).

Figure 3.

Figure 3. Mass as a function of spectral type for objects with directly measured individual masses in our sample (filled red circles) and from the literature (filled black diamonds). Open red circles show objects from our sample with individual masses computed from our measured total mass and SM08 model-derived mass ratios. The likely unresolved multiple systems (gray circles) have high masses for their spectral type, as expected. Otherwise, no objects later than L4 have masses above 70 ${M}_{\mathrm{Jup}}$, implying that this is roughly the spectral type and mass corresponding to the hydrogen-fusion limit for the field population. The lowest mass object in our sample (SDSS J0423−0414B) does not have the latest spectral type, illustrating that our field sample spans a range of ages. (To clearly view many data points at the same spectral type, small offsets to the spectral type values have been added for plotting here.)

Standard image High-resolution image

Figure 4 shows our dynamical mass sample on the color–magnitude diagram (CMD) with data points colored by the measured masses. Overall, mass broadly decreases through the CMD sequence, although there is no one-to-one correspondence between mass and CMD location, as expected for our sample that is drawn from the field population of ultracool dwarfs spanning a wide range of ages. The clearest illustration of this is that the lowest mass objects happen to be located roughly in the middle of the L/T transition, while the latest-type, bluest objects are somewhat more massive. This more likely reflects the distribution of ages in our sample rather than indicating that later-type objects tend to be more massive. Figure 5 shows our measured absolute magnitudes as a function of spectral type, colored according to mass, in order to assess the reliability of both plotted quantities. Our sample is broadly consistent with the field sequence, indicating the accuracy of our spectral types, with the largest outlier being 2MASS J0920+3517B, which is consistent with our dynamical masses showing that this is an unresolved pair of brown dwarfs in a triple system.

Figure 4.

Figure 4. Color–magnitude diagram (CMD) showing our dynamical mass sample with symbols colored according to their directly measured individual masses. (The color bar is the same as that used in all other plots with this color scheme.) Upward-pointing triangles indicate primary components, downward-pointing triangles indicate secondary components, and the field sequence is shown for reference as gray squares using the latest compilation of ultracool dwarf parallaxes at http://www.as.utexas.edu/~tdupuy/plx. (We plot only normal, single, field objects that have S/N > 10 parallaxes and JK uncertainties < 0.10 mag.) Although mass generally decreases through the CMD sequence, there is no one-to-one correspondence with mass and CMD location given that our sample is drawn from the field population of ultracool dwarfs spanning a range of ages. For example, the latest-type T dwarfs ($J-K=-0.5$ to 0.0 mag) happen to not be the least massive objects in our sample. By chance, the least massive objects are all located roughly in the middle of the L/T transition ($J-K=0.5\mbox{--}1.0$ mag). The massive object in the bottom-right part of the CMD (black downward-pointing triangle near MJ = 14 mag) is 2MASS J0920+3517B, an unresolved pair of brown dwarfs in a triple system.

Standard image High-resolution image
Figure 5.

Figure 5. Absolute magnitude plotted as a function of spectral type for our dynamical mass sample with symbols colored according to their directly measured individual masses. (The color bar is the same as that used in all other plots with this color scheme.) Upward-pointing triangles indicate primary components, downward-pointing triangles indicate secondary components, and the field sequence is shown for reference as gray squares using the latest compilation of ultracool dwarf parallaxes at http://www.as.utexas.edu/~tdupuy/plx. (We plot only normal, single, field objects that have S/N > 10 parallaxes and JK uncertainties < 0.10 mag.) Our sample is broadly consistent with the field sequence, indicating the accuracy of our spectral types that we derived by decomposing each binary's integrated-light spectrum using our resolved photometry and a library of spectral templates. The largest outlier is 2MASS J0920+3517B (L9.0 ± 1.5 sitting well above the L/T transition sequence on the J-band plot), which is consistent with our dynamical masses showing that this is an unresolved pair of brown dwarfs in a triple system. 2MASS J0700+3157B, the other such unresolved pair, is not an outlier on these plots because our derived spectral type (L6.5 ± 1.5) and absolute magnitudes happen to be consistent with the field sequence.

Standard image High-resolution image

To provide a simple summary of our results, Table 12 reports the mean mass in seven spectral type bins selected to contain three to seven objects per bin, excluding unusual objects (unresolved binaries and the pre-main-sequence system LP 349-25AB). As expected, the mean mass broadly decreases with spectral type. We caution that this table is simply meant to provide a guide to the typical masses of objects in the field population and is not intended to be used for any quantitative astrophysical purpose. At earlier spectral types, Table 12 may be useful for providing model-independent mass estimates for very low-mass stars. Next, we briefly compare these results to previous work and defer discussion of the astrophysical interpretation of our individual masses to later sections.

Table 12.  Average Mass Per Spectral Type Bin

Spectral Mass (${M}_{\mathrm{Jup}}$) Number
Type mean rms in bin
M7–M7.5 102 8 3
M8–M8.5 94 11 7
M9–L0.5 81 10 7
L1–L3.5 76 6 4
L4–L7.5 58 9 7
L8–T1.5 48 14 3
T2–T5.5 36 9 5

Note. The mean is computed as the weighted average among the individual mass measurements in each spectral type bin. Bin sizes vary in order to include a reasonable number of objects per bin, and the bins are larger at later types. From our own 38 individual masses we exclude the pre-main-sequence binary LP 349-25AB, objects that are possible unresolved binaries (LP 415-20A, 2MASS J0700+3157B, and 2MASS J0920+3517B), and LHS 2397aB (66 ± 4 ${M}_{\mathrm{Jup}}$) because it has no directly measured spectral type. Individual masses from the literature that we include here are those for LHS 1070BC (M9.5+L0; Köhler et al. 2012) and Gl 569Bab (M8.5+M9; total mass from this work and mass ratio from Konopacky et al. 2010).

Download table as:  ASCIITypeset image

5.2.1. Comparison to Literature Mass Ratios

Four of our 19 systems with individual masses have mass ratios derived from radial velocities by Konopacky et al. (2010). They quote mass ratios as M1/M2, rather than the standard q ≡ M2/M1, so for the purposes of comparison here we quote our measurements as 1/q as well. For 2MASS J0746+2000AB we find 1/q = 1.05 ± 0.03, which agrees within the large uncertainties of ${4.0}_{-3.8}^{+0.1}$ as found by Konopacky et al. (2010). They report a similar mass ratio for 2MASS J2140+1625AB with much smaller uncertainties (${4.0}_{-0.1}^{+0.0}$), which disagrees at ≈8σ with our result (${1.65}_{-0.23}^{+0.19}$). This seems to confirm the suggestion in Section 5.4.5 of Konopacky et al. (2010) that their mass ratio for this binary was likely more uncertain than the quoted errors implied. One of our most precise mass ratios is for LHS 2397aAB ($1/q={1.42}_{-0.06}^{+0.05}$), and it agrees well with the value of ${1.5}_{-1.4}^{+7.1}$ reported by Konopacky et al. (2010). Finally, we find 1/q = 1.06 ± 0.03 for LP 349–25AB, which disagrees with the value of 0.5 ± 0.3 found by Konopacky et al. (2010). As discussed in Section 4.2 of Dupuy et al. (2010), a mass ratio that agrees with our value can be accommodated by the radial velocities of Konopacky et al. (2010), suggesting that their quoted errors may be underestimated.

The only other directly measured mass ratio in the literature for a system in our sample is from Harris et al. (2015) for 2MASS J0746+2000AB. Their USNO astrometry gives a value of q = 0.925, which agrees well with our value of ${0.952}_{-0.027}^{+0.026}$.

5.3. Bolometric Luminosities

We have derived luminosities for all objects in our sample using our resolved photometry along with the empirical relations between luminosity and absolute magnitude that are described in Appendix A.3. This is somewhat different from our past work where we relied on bolometric correction–spectral type relations to derive individual luminosities. Our new approach obviates the need to reference spectral types, as it links absolute magnitude directly to luminosity. Given that J- and H-band absolute magnitudes get brighter and plateau across the L/T transition, respectively, only K-band absolute magnitudes are suitable for the entire range of luminosity of our sample. The luminosity scatter about the K-band polynomial relation is somewhat higher across the L/T transition (0.07 dex; MK > 13.0 mag) than at brighter magnitudes (0.04 dex). Interestingly, if we only consider the late-M and earlier L dwarfs with MH < 13.3 mag, the scatter in luminosity about the H-band relation is significantly lower (0.023 dex). It is not obvious why this should be the case from an astrophysical perspective, but we nonetheless use this to our advantage. For binaries where both components have MH < 13.3 mag, we use the H-band relation with luminosity, and for the remainder we use the K-band relation. In all cases we used whichever photometry was more precise between 2MASS and MKO. We report the resulting component luminosities in Table 11, and Figure 6 shows our derived luminosities as a function of our individual masses.

Figure 6.

Figure 6. Luminosity as a function of mass for objects with directly measured individual masses in our sample (filled red circles) and from the literature (filled black diamonds). Open red circles show objects from our sample with individual masses computed from our measured total mass and SM08 model-derived mass ratios. Filled gray circles indicate the likely unresolved multiples in our sample 2MASS J0920+3517B and 2MASS J0700+3157B. The other possible unresolved multiple LP 415-20A (${156}_{-18}^{+17}$ ${M}_{\mathrm{Jup}}$) does not appear on this plot because it is too massive. Cond model isochrones are shown for reference, indicating that most objects in our sample are consistent with having ages of ∼1–10 Gyr. For Gl 569Bab we use black diamonds to indicate that the literature mass ratio is used to compute individual masses, even though it is in our sample and we use our dynamical total mass here.

Standard image High-resolution image

As a check on our method, we compare our results to published work using integrated-light spectra and photometry to derive luminosities. For objects in common with Golimowski et al. (2004a), Dieterich et al. (2014), and Filippazzo et al. (2015), all of whom use somewhat different methods and generations of models for deriving luminosities, we find that our total luminosities agree well within the errors after accounting for the different distances assumed in each work. For example, summing our luminosities for the components of 2MASS J0746+2000AB, we find $\mathrm{log}({L}_{\mathrm{bol}}/{L}_{\odot })=-3.375\pm 0.020$ dex, which agrees well with published values of −3.41 ± 0.02 dex (Golimowski et al. 2004a), −3.413 ± 0.009 dex (Dieterich et al. 2014), and −3.391 ± 0.003 dex (Filippazzo et al. 2015). Other objects in common include Kelu-1, LHS 2397a, SDSS J0423−0414, and 2MASS J1728+3948.

6. Evolutionary Model-derived Properties

Directly measured masses and luminosities are just two of the many physical properties that are of interest in characterizing our sample of very low-mass stellar and brown dwarf binaries. In principle, all physical properties of both stars and brown dwarfs can be predicted from just a few fundamental parameters: mass, age, and composition. (Other properties such as the entropy of formation, initial angular momentum, and magnetic fields may also be important as initial conditions.) Our dynamical mass sample consists of binaries that may be conservatively presumed to be coeval to within a few Myr and to be composed of the same material. It is therefore an ideal sample for pairing with evolutionary models, since the three most fundamental parameters are all constrained at some level.

We consider two families of evolutionary models here, the Lyon group's models and the SM08) models (Figure 7 shows all model grids used here). The most recent grid from the Lyon group comes from Baraffe et al. (2015), hereinafter BHAC, and extends over masses of 0.01–1.4 M. The published BHAC grid samples masses in the range 0.02–0.1 M in increments of 0.01 M (with two extra models at 0.072 M and 0.075 M) and only tracks evolution down to ${T}_{\mathrm{eff}}=1500$ K. We have obtained BHAC model tracks that extend down to 1300 K (only for fundamental properties, not absolute magnitudes), as well as more finely gridded tracks over the mass range 0.05–0.07 M in 0.001 M increments. This enhanced BHAC evolutionary model grid was kindly provided by I. Baraffe (2016, private communication) so that we could more accurately sample the hydrogen-fusion and lithium-fusion mass limits. Many of our sample binary components have luminosities too low to be covered by the BHAC grid, and when this is the case we use the Lyon Cond models instead (Baraffe et al. 2003). We do not consider the Lyon Dusty models (Chabrier & Baraffe 2000) as these have been supplanted by BHAC models over the luminosity range for which they would be appropriate (namely late-M and L dwarfs). The second family of models we consider is from Saumon & Marley (2008), hereinafter SM08, which in contrast to BHAC only covers cooler temperatures (${T}_{\mathrm{eff}}\lesssim 2400$ K; $\mathrm{log}({L}_{\mathrm{bol}}/{L}_{\odot })\lesssim -3.3$ dex) and thus many of our more luminous objects are not covered by the SM08 grid. These hard limits on the model grids mean that we cannot always report model-derived properties for a given component because we do not extrapolate beyond the range of any model grid.

Figure 7.

Figure 7. 

Evolutionary models used in our analysis. Each colored line shows the predicted luminosity as a function of age for an object of a given mass. The 75 ${M}_{\mathrm{Jup}}$, 70 ${M}_{\mathrm{Jup}}$, and 60 ${M}_{\mathrm{Jup}}$ models are plotted with thicker lines as a visual aid. All model masses are integer multiples of 0.001 M (1.048 ${M}_{\mathrm{Jup}}$), and only masses of 0.015–0.100 M (15–105 ${M}_{\mathrm{Jup}}$) are plotted here. BHAC models are uniformly sampled at steps of 0.001 M over the range 0.050–0.070 M (52–73 ${M}_{\mathrm{Jup}}$) and at coarser steps at other masses. Dotted and dashed lines indicate boundaries in lithium depletion of 50% and 99.9% depleted, respectively. BHAC models are only available down to ${T}_{\mathrm{eff}}\approx 1300$ K, while SM08 models are only available below ≈2400 K. (The color bar is the same as that used in all other plots with this color scheme.) (An extended version of this figure is available.)

Standard image High-resolution image

    In order to perform tests of the models, we first use the observed properties of mass and luminosity to infer other properties from evolutionary models. For our entire sample we have individual luminosities and precise total masses. In most cases (19 of 22 systems) we also have individual masses, but the precision of our mass ratios causes most of these individual masses to be less precise than the total mass. Therefore we take a two-pronged approach to inferring physical properties from models, either using the more precise total mass alone or the full individual mass information.

    The total-mass analysis we use here is distinct from methods that we have used in our previous works (e.g., Liu et al. 2008; Dupuy et al. 2009b) as well as other approaches in the brown dwarf literature (see Section 4.5 of Dupuy et al. 2010 for a discussion of different methods). Our past Monte Carlo methods relied on a two-step interpolation, first using the individual luminosities to compute a model-predicted total mass as a function of age, and then applying our observational constraint on the total mass to determine the age distribution. Such a method is susceptible to numerical problems where more than one mass corresponds to a given luminosity either at a given age (e.g., due to deuterium burning) or over a range of ages (e.g., even very low-mass stars become slightly more luminous at old ages than they were on the zero-age main sequence). More importantly, measurement errors can cause Monte Carlo samples to fall in regions of parameter space not covered by models, e.g., a luminosity that scatters low so that models predict a star of that mass would never be that faint, which causes a simple interpolation approach to fail. Such interpolation problems can also occur in the case of individual-mass analysis methods applied to stars, even though they have worked well for brown dwarfs and young stars in our previous studies (Dupuy et al. 2015b, 2016).

    Therefore, we have developed a new approach for deriving physical properties from models that is based on the statistical technique of rejection sampling. This approach works well when the number of parameters of interest is small, and in the individual-mass analysis we only need to find the probability distribution for one unknown parameter—age—because age combined with mass allows all other properties to be directly interpolated from models. Our approach begins with a uniform distribution of Monte Carlo-drawn ages spanning the minimum to maximum ages of a given model grid, and through rejection sampling we end up with a distribution of ages that match the observed luminosities at the measured masses. For the case of individual masses this is conceptually straightforward, but when using only the total mass we must simultaneously try both random ages and component masses (i.e., essentially allowing for one more unknown parameter, the mass ratio). In both cases, we use the final individual mass and age samples remaining at the end of the rejection sampling to derive other properties from models (${L}_{\mathrm{bol}}$, radius, ${T}_{\mathrm{eff}}$, etc.). We now describe each case in more detail.

    In our individual-mass analysis, we begin with 106 Monte Carlo trials of age that are randomly paired with masses from our MCMC results. The input distribution of ages defines the prior on age, and we choose uniformly distributed values, i.e., a flat prior. Each mass and age uniquely corresponds to a luminosity from models, which we determine from bilinear interpolation of a uniform 2D grid of mass and age constructed using Delaunay triangulation as implemented in the trigrid function in IDL. We compute a χ2 for each sample given our measured median ${L}_{\mathrm{bol}}$ and its uncertainty. We account for the fact that our luminosity and mass measurements are correlated (due to the common parallax) by fitting a line to $\mathrm{log}({L}_{\mathrm{bol}})$ as a function of $\mathrm{log}(M)$. At each Monte Carlo-drawn mass, the effective luminosity (${L}_{\mathrm{bol}}^{\prime }$) is given by the coefficients of this linear fit. The observational error in luminosity at fixed mass (${\sigma }_{{L}_{\mathrm{bol}}^{\prime }}$) is given by the rms scatter about the fit, which is equal to or smaller than the actual ${L}_{\mathrm{bol}}$ error depending on the degree of covariance. Thus, we compute ${\chi }^{2}={({L}_{\mathrm{bol},\mathrm{model}}-{L}_{\mathrm{bol}}^{\prime })}^{2}/{\sigma }_{{L}_{\mathrm{bol}}^{\prime }}^{2}$ and a corresponding probability of $p={e}^{-({\chi }^{2}-\min ({\chi }^{2}))/2}$, normalized by the Monte Carlo sample with the lowest χ2. We then draw random, uniformly distributed variates u, and we rejected samples where p < u. In order to efficiently sample the most relevant ages, we iterated this process gradually, narrowing the trial age distribution based on the results of the previous iteration. We found that after three iterations the number of successful Monte Carlo trials stabilized. After determining our final age distribution, we then determined other properties (${T}_{\mathrm{eff}}$, lithium depletion, etc.) from 2D grids constructed with trigrid in the same way as for luminosity.

    In our total-mass analysis we have two unknown parameters, age and mass ratio. Therefore, we not only draw random, uniformly distributed ages as above, but also uniformly distributed values of $\mathrm{log}({M}_{1})$ for one component. The mass of the other component can then be computed from the total mass in that Monte Carlo sample, ${M}_{2}={M}_{\mathrm{tot}}-{M}_{1}$. We initially sample all ages covered by the model tracks and masses for M1 ranging from the lowest mass model to the maximum total mass in the MCMC chain. We interpolate the luminosity from models for each component mass and age pair and compute a total χ2 by comparing the interpolated luminosities for both components to our measurements. As in our individual-mass analysis, we account for correlation between our measured total mass and component luminosities by fitting lines to $\mathrm{log}({L}_{\mathrm{bol},1})$ and $\mathrm{log}({L}_{\mathrm{bol},2})$ as a function of $\mathrm{log}({M}_{\mathrm{tot}})$. We reject Monte Carlo samples based on the probabilities computed from the sum of the components' χ2 and then iteratively narrow the range of age and component mass searched, again finding three iterations sufficient in all cases. We use the final remaining samples of age and component masses to interpolate other properties (${T}_{\mathrm{eff}}$, etc.) from the 2D model grids.

    Table 13 show the results of our analyses using total mass and individual luminosities as well as individual masses and luminosities, when possible. Our rejection sampling analysis naturally produces output distributions of individual luminosities and masses in all cases. Conceptually, this is because any inputs to the rejection sampling (even measured quantities) can be thought of as priors on those properties, so the output distributions may be somewhat different (narrower or slightly shifted). The final output properties will always be fully consistent with model predictions. For example, in the case of a star on the main sequence, combining our measured luminosity and age prior typically results in a much narrower range of masses according to models because only certain masses are predicted to correspond to the input luminosity. In most cases, the median mass and luminosity output by the rejection sampling analysis agrees within ≲1.5σ of the input values, as expected, but exceptional cases are discussed in the following sections. Figure 8 shows our input measurements compared to model tracks and the resulting age distributions from both individual-mass and total-mass analyses.

    Figure 8.

    Figure 8.

    Left panel: total mass derived from models using the component luminosities over a range of ages (purple), where the vertical extent of the shaded region corresponds to the 1σ uncertainties in the luminosities. Our measured total mass and 1σ confidence interval are shown by a solid line and by dashed lines, respectively. (This panel is for display purposes only as our actual analysis is based on the rejection sampling as described in Section 6.) Middle panel: our directly measured individual masses and luminosities are plotted as black filled circles with 1σ error bars. Model isochrones are shown as gray lines. The 1σ interval in mass and luminosity after performing our rejection sampling analysis is shown by the colored error bars for the primary (blue) and secondary (red). When there are offsets between these colored points and the measurements, or differences in the error bars, it is caused by rejecting input Monte Carlo samples that do not agree with models and/or by imposing our uniform age prior. Right panel: final age distribution derived from our rejection sampling analysis using the total mass and individual luminosities (solid purple), the primary mass and luminosity (blue hatched), and the secondary mass and luminosity (red hatched). Our uniform prior in age results in many stellar binaries having nearly flat age distributions over the range of main-sequence ages covered by models, whereas brown dwarfs and pre-main-sequence stars like LP 349-25AB have well-constrained model-derived ages. The figure set contains all the figures. (The complete figure set (25 images) is available.)

    Standard image High-resolution image

    Table 13.  Observed and Model-derived Properties

      Using Total Mass Using Individual Masses
    Property Primary Secondary ${\rm{\Delta }}=B-A$ Primary Secondary ${\rm{\Delta }}=B-A$
    Input Observed Properties for LP 349-25AB
    Mass M (${M}_{\mathrm{Jup}}$) ${166}_{-7}^{+6}$ 85 ± 4 80 ± 3 $-{5.0}_{-2.7}^{+2.6}$
    Mass ratio q ${0.941}_{-0.030}^{+0.029}$
    log(${L}_{\mathrm{bol}}$) [${L}_{\odot }$] −3.075 ± 0.026 −3.198 ± 0.027 $-{0.12}_{-0.04}^{+0.03}$
    Derived from Baraffe et al. (2015) Evolutionary Models for LP 349-25AB
    Mass M (${M}_{\mathrm{Jup}}$) 89 ± 4 ${78}_{-4}^{+3}$ −10 ± 3 86 ± 4 81 ± 3 −5 ± 5
    log(${L}_{\mathrm{bol}}$) [${L}_{\odot }$] $-{3.074}_{-0.032}^{+0.029}$ −3.20 ± 0.03 −0.12 ± 0.04 $-{3.076}_{-0.029}^{+0.031}$ $-{3.199}_{-0.031}^{+0.029}$ −0.12 ± 0.04
    Mass ratio q ${0.88}_{-0.04}^{+0.03}$ ${0.94}_{-0.06}^{+0.05}$
    Age t (Gyr) ${0.271}_{-0.029}^{+0.022}$ ${0.251}_{-0.033}^{+0.026}$ ${0.30}_{-0.04}^{+0.03}$ 0.05 ± 0.05
    $\mathrm{log}(t)$ [years] 8.43 ± 0.04 8.40 ± 0.05 8.47 ± 0.05 0.07 ± 0.07
    ${T}_{\mathrm{eff}}$ (K) ${2740}_{-29}^{+32}$ 2620 ± 30 −120 ± 40 ${2729}_{-27}^{+26}$ ${2629}_{-27}^{+29}$ −100 ± 40
    Radius (${R}_{\mathrm{Jup}}$) ${1.255}_{-0.024}^{+0.019}$ ${1.193}_{-0.017}^{+0.020}$ −0.061 ± 0.021 ${1.262}_{-0.026}^{+0.025}$ ${1.181}_{-0.023}^{+0.021}$ −0.08 ± 0.03
    $\mathrm{log}(g)$ [cm s−2] ${5.143}_{-0.020}^{+0.024}$ 5.133 ± 0.024 $-{0.010}_{-0.004}^{+0.005}$ ${5.125}_{-0.029}^{+0.027}$ 5.156 ± 0.027 0.03 ± 0.04
    log(Li/Liinit) <−4.0 <−4.0 <−4.0 <−4.0
    MKO(JK) (mag) 0.808 ± 0.002 ${0.810}_{-0.003}^{+0.005}$ ${0.003}_{-0.002}^{+0.003}$ ${0.809}_{-0.003}^{+0.004}$ ${0.806}_{-0.004}^{+0.003}$ −0.003 ± 0.005
    MKO(JH) (mag) 0.492 ± 0.004 ${0.486}_{-0.002}^{+0.001}$ −0.006 ± 0.004 ${0.491}_{-0.003}^{+0.004}$ 0.487 ± 0.002 −0.005 ± 0.004

    Only a portion of this table is shown here to demonstrate its form and content. A machine-readable version of the full table is available.

    Download table as:  DataTypeset image

    Our new approach to deriving properties from models via rejection sampling gives essentially identical results as our past approach for cases where a given mass and luminosity measurement lie well within the bounds of model isochrones, i.e., for brown dwarfs and young stars. The main advantage to rejection sampling is that it properly handles stars on the main sequence and that it explicitly specifies a prior on age. For example, in our past work on LHS 1901AB, we quoted median and ±1σ values from our age posterior from Lyon models of ${0.28}_{-0.08}^{+9.72}$ Gyr (Dupuy et al. 2010). This low median age was a consequence of directly interpolating our measured luminosity that is slightly higher than, but still well within, the 1σ errors of the model-predicted main-sequence luminosity for objects of that mass. Using our new method, we report a more sensible result of ${5.2}_{-4.7}^{+1.9}$ Gyr, thanks to both allowing measured luminosities to scatter above and below the main-sequence value and imposing an explicit age prior. In addition, in our new approach we always interpolate directly from model grids in mass and age, which circumvents issues related to the fact that luminosity can be double-valued at a given mass as a function of age.

    Finally, we note that the models used in our analysis all assume solar metallicity. Different assumptions about metallicity would impact opacities and likely also influence cloud formation and evolution, both of which would change how luminosity evolves with time. In our work on Gl 417BC (Dupuy et al. 2014), we discussed this impact of changing opacities on cooling, inferring from the cloudless SM08 models that a change of ±0.3 dex in metallicity changes luminosity by no more than ±0.05 dex (super-solar models are more luminous at a given mass and age). Thus, if our sample is on average significantly offset from solar metallicity for some reason, this would result in only a slight systematic shift in our results. For reference, Santos et al. (2008) report a mean and rms metallicity of −0.10 ± 0.24 dex for stars in the solar neighborhood. Even for such relatively large metallicity differences as ±0.3 dex, the corresponding difference in luminosity is comparable to our typical luminosity uncertainties, so we do not expect metallicity to strongly affect our conclusions.

    7. Discussion

    Our large sample of dynamical masses enables a broad array of empirical tests of substellar evolution, many of which have not been possible before without precise individual masses. We now discuss a number of key topics, both in testing model predictions and establishing empirical relations. In the following analysis we will mostly deal with our results as a whole, but a detailed discussion of each individual system is provided in Appendix B. When discussing our results below, we generally do not explicitly exclude anomalous objects (e.g., the unresolved components of triple systems discussed in Section 7.4) every time they might be relevant.

    7.1. Empirical Test of the Substellar Boundary

    Our work represents the first sample of individual masses for spectrally classified L and T dwarfs. This allows us to examine the maximum mass of the latest-type objects, which is the mass of the boundary between stars and brown dwarfs. Figure 6 shows our 38 individual mass and luminosity measurements along with all other published model-independent masses, including the spectrally unclassified objects Gl 802B (80 ± 14 ${M}_{\mathrm{Jup}}$; Ireland et al. 2008) and HR 7672B (${68.7}_{-3.1}^{+2.4}$ ${M}_{\mathrm{Jup}}$; Crepp et al. 2012), as well as the M8.5+M9 binary Gl 569Bab (${80}_{-8}^{+9}$ ${M}_{\mathrm{Jup}}$ and ${58}_{-9}^{+7}$ ${M}_{\mathrm{Jup}}$; Konopacky et al. 2010), the M9.5+L0 binary LHS 1070BC (81 ± 5 ${M}_{\mathrm{Jup}}$ and 74 ± 4 ${M}_{\mathrm{Jup}}$; Köhler et al. 2012), and the M7+M7 binary LSPM J1314+1320AB (92.8 ± 0.6 ${M}_{\mathrm{Jup}}$ and 91.7 ± 1.0 ${M}_{\mathrm{Jup}}$; Dupuy et al. 2016). None of the components violate the most basic predictions of models within the observational uncertainties, except for systems suspected to be higher-order multiples based on our analysis independent of models. No objects have a lower luminosity than expected given their mass and the finite age of the universe.

    Figure 3 shows our individual masses as a function of spectral type. The lack of objects at high mass and late spectral type is empirical evidence for a mass limit to hydrogen fusion. Moreover, examining the maximum mass of the latest-type objects (or equivalently the lowest luminosity objects in Figure 6) allows us to constrain the upper limit on the mass of the substellar boundary, and correspondingly a lower limit on the mass of main-sequence stars. In fact, our dynamical mass sample is (somewhat unfortunately) ideal for this test because we are biased toward high masses. The main observational limitation in measuring masses to date has been achieving the time baseline needed for robust orbit determinations of binaries with typical periods of 20–30 years. The highest mass late-L or T dwarfs in our sample are 2MASS J1728+3948B (L7.0 ± 1.0; 67 ±5 ${M}_{\mathrm{Jup}}$; $\mathrm{log}({L}_{\mathrm{bol}}/{L}_{\odot })=-4.49\pm 0.04$ dex) and 2MASS J1404−3159A (L9.0 ± 1.0; 65 ± 6 ${M}_{\mathrm{Jup}}$; $\mathrm{log}({L}_{\mathrm{bol}}/{L}_{\odot })=-4.52\,\pm 0.05$ dex). The next most massive late-type object is 2MASS J2132+1341B (L8.5 ± 1.5; 60 ± 4 ${M}_{\mathrm{Jup}}$; $\mathrm{log}({L}_{\mathrm{bol}}\,/{L}_{\odot })=-{4.50}_{-0.04}^{+0.05}$ dex). Going to earlier types, there are three objects that have masses and spectral types that are consistent among each other within the errors: 2MASS J0920+3517A (L5.5 ± 1.0; 71 ± 5 ${M}_{\mathrm{Jup}}$; $\mathrm{log}({L}_{\mathrm{bol}}/{L}_{\odot })=-4.28\pm 0.03$ dex), 2MASS J1728+3948A (L5.0 ± 1.0; 73 ± 7 ${M}_{\mathrm{Jup}}$; $\mathrm{log}({L}_{\mathrm{bol}}\,/{L}_{\odot })=-{4.29}_{-0.05}^{+0.04}$ dex), and 2MASS J2132+1341A (L4.5 ± 1.5; 68 ± 4 ${M}_{\mathrm{Jup}}$; $\mathrm{log}({L}_{\mathrm{bol}}/{L}_{\odot })=-4.22\pm 0.05$ dex). Despite the fact that we are more sensitive to massive systems, even our most massive late-type objects are only consistent with having masses as high as 71 ${M}_{\mathrm{Jup}}$ within 1σ, and even at spectral types as early as L4, the most massive objects are consistent with 70 ${M}_{\mathrm{Jup}}$. Therefore, we estimate an empirical substellar boundary of ≈70 ${M}_{\mathrm{Jup}}$.

    The theoretical mass limit of hydrogen fusion quoted in the literature varies widely. One review by Burrows et al. (2001) places the minimum mass for H fusion between 0.070–0.075 M (73–79 ${M}_{\mathrm{Jup}}$) for solar metallicity and at 0.092 M (96 ${M}_{\mathrm{Jup}}$) for zero metallicity. More recent models from Burrows et al. (2011) also show a range of 0.070–0.075 M for the H-fusion mass assuming different helium fractions (their Figure 6). A review by Chabrier & Baraffe (2000) gives the minimum mass for H fusion as 0.070 M (73 ${M}_{\mathrm{Jup}}$) when cloud opacity is included and 0.072 M (75 ${M}_{\mathrm{Jup}}$) for cloudless atmospheres. If the H-fusion mass limit is indeed as high as ≈75 ${M}_{\mathrm{Jup}}$, then it is surprising that we have not uncovered objects with masses closer to this limit in our sample. Figure 7 shows more recent tracks from BHAC that includes a 0.072 M (75 ${M}_{\mathrm{Jup}}$) track that stabilizes in luminosity after ≈2 Gyr and thus we would consider a star. Some of the next lower mass tracks appear more star-like in that sense, but by 0.067 M (70 ${M}_{\mathrm{Jup}}$) the luminosity appears to be on course to monotonically decrease with the age of the universe. Finally, SM08 quote a minimum mass of 0.070 M (73 ${M}_{\mathrm{Jup}}$) from their cloudy models, because that track ultimately reaches 1550 K where clouds are very significant in the atmosphere. Thus, the lack of any ≳70 ${M}_{\mathrm{Jup}}$ brown dwarfs in our sample is consistent with the lower range of predictions in the literature for the minimum mass for H fusion (70–73 ${M}_{\mathrm{Jup}}$).

    If we adopt our empirical substellar boundary of ≈70 ${M}_{\mathrm{Jup}}$, we can then determine the latest spectral type of objects above this boundary (i.e., objects that may be stars at the very bottom of the main sequence). The most precise stellar mass in our sample is for 2MASS J0746+2000B (L1.5 ± 0.5; 78.4 ± 1.4 ${M}_{\mathrm{Jup}}$; $\mathrm{log}({L}_{\mathrm{bol}}/{L}_{\odot })=-{3.777}_{-0.027}^{+0.028}$ dex), but this is likely well above the H-fusion limit. Going to later spectral types we find 2MASS J0700+3157A (L3.0 ± 1.0; 68.0 ± 2.6 ${M}_{\mathrm{Jup}}$; $\mathrm{log}({L}_{\mathrm{bol}}/{L}_{\odot })=-3.95\pm 0.04$ dex) and 2MASS J1017+1308B (L3.0 ± 1.0; 75 ± 7 ${M}_{\mathrm{Jup}}$; $\mathrm{log}({L}_{\mathrm{bol}}/{L}_{\odot })=-3.84\pm 0.04$ dex). After this is the cluster of ≈L4–L6 objects with masses consistent with 70 ${M}_{\mathrm{Jup}}$ mentioned above. We therefore conclude that the end of the main sequence occurs within the range of spectral types L3–L5 and luminosities of 10−4.3 to 10−3.9 dex, independent of model assumptions. Objects with spectral types later than this, or luminosities lower than this, are brown dwarfs, while objects with earlier types or higher luminosities could either be stars or brown dwarfs. Our result is consistent with early works on the first samples of L dwarfs that showed the substellar boundary was likely within this spectral type range (e.g., Kirkpatrick et al. 1999). A larger sample and more precise component spectral types would enable us to refine the location of this empirically defined substellar boundary.

    Finally, we discuss our results in the context of the work of Dieterich et al. (2014) on the substellar boundary. Their approach uses the minimum radius (determined via luminosity and a temperature derived from model atmospheres) on the Hertzsprung–Russell diagram. The physical motivation is that the transition to degeneracy-dominated interiors marks the beginning of the substellar regime, as degeneracy pressure is what prevents objects below a critical mass from reaching sufficiently high core temperatures to generate a significant amount of energy by fusing hydrogen. To perform their test, they start at the late-M dwarfs and examine progressively cooler objects that have progressively smaller radii, until the trend reverses and cooler objects no longer have significantly smaller radii. Dieterich et al. (2014) find that this occurs at a radius of 0.086 ± 0.003 ${R}_{\odot }$ for the object 2MASS J0523−1403, which has a spectral type of L2.5 and $\mathrm{log}(L/{L}_{\odot })=-3.9$ dex, and they conclude that this marks the end of the main sequence. Compared to the boundary derived here using mass, their result is somewhat earlier in type and more luminous than our result but broadly consistent within the uncertainties. Dieterich et al. (2014) discuss the fact that models predict a locus for their radius test at somewhat cooler temperatures and suggest that lowering the metal abundances adopted in current models could explain the discrepancy; however, they also note that such a change in the abundances would alter the mass limit of the boundary to higher masses. The fact that we find a rather low mass limit for the substellar boundary would therefore seem to indicate that abundances alone will not rectify the tensions between models and the results of Dieterich et al. (2014). We speculate that one possibility could be systematic errors in temperatures derived from model atmospheres akin to those identified in our previous work on late-M dwarfs (Dupuy et al. 2010), where model atmospheres gave higher temperatures than implied by evolutionary model radii and measured luminosities.

    7.2. Lithium Fusion Mass Limit

    Theoretical predictions of lithium depletion in ultracool dwarfs as a function of mass and age are thought to be one of the most reliable methods for age-dating young clusters (e.g., Chabrier et al. 1996; Bildsten et al. 1997; Binks & Jeffries 2014; Kraus et al. 2014). They have also long been used as a method of confirming ultracool dwarfs as substellar (e.g., Rebolo et al. 1992; Basri et al. 1996). The brown dwarfs in our sample span masses both above and below the predicted lithium depletion boundary, and many of them are in systems with published optical, integrated-light spectroscopy that constrains the presence or absence of lithium. One potential complication is that the Li i doublet at 6708 Å, which provides the most sensitive observational probe of lithium, is expected to disappear in cool objects where monatomic lithium becomes incorporated into LiCl and LiOH molecules or at cooler temperatures condenses into LiF and Li2S (e.g., Lodders 1999). This chemical depletion at low temperatures has not previously been tested with objects of known mass. The latest-type object known to display Li i absorption is WISE J1049−5319B (T0; Faherty et al. 2014; Lodieu et al. 2015), implying that any of the L dwarfs in our sample might plausibly display Li i absorption if they are not depleted by Li fusion.

    Figure 9 shows our individual masses as a function of age for the 13 systems with well-constrained model-derived ages, i.e., systems containing at least one brown dwarf, as well as the pre-main-sequence binary LP 349–25AB. (The low-mass main-sequence binaries in our sample have unconstrained ages, and as expected, none are known to display Li i absorption.) Only 2 of these 13 binaries have published lithium detections, SDSS J0423−0414AB (EW = 11 Å; Kirkpatrick et al. 2008) and Gl 417BC (EW = 11.5 Å; Kirkpatrick et al. 2001). Seven more systems have published spectra that exclude lithium absorption in integrated light: LP 349–25AB (<0.5 Å; Reiners & Basri 2009), 2MASS J0700+3157AB (<0.3 Å; Thorstensen & Kirkpatrick 2003), LHS 2397aAB (<0.5 Å; Reiners & Basri 2009), 2MASS J1728+3948AB (<4 Å; Kirkpatrick et al. 2000), 2MASS J2132+1341AB (Cruz et al. 2007), and DENIS J2252−1730AB (Reid et al. 2008b). (See Appendix B for a detailed discussion of all of these binaries and their lithium constraints.)

    Figure 9.

    Figure 9. Individual masses plotted at their system ages for 13 of our sample binaries that have well-constrained ages according to evolutionary models. Open symbols indicate systems where we use model-derived mass ratios to compute individual masses, including Gl 569Bab. Upward-pointing blue triangles indicate components of binaries where lithium is detected in integrated light, while downward-pointing orange triangles indicate systems for which an optical spectrum is available and does not display lithium absorption. Model-predicted mass limits for strong (>99.9%) and weak (<10%) lithium depletion are shown as thick black and thin gray lines, respectively. The models shown here are BHAC (Baraffe et al. 2015) and Tucson (Burrows et al. 1997). Lithium detections and nondetections seem to generally agree well with BHAC model-predicted lithium depletion, within the observational uncertainties on mass. The mass limit for lithium depletion is higher in Tucson models (70 ${M}_{\mathrm{Jup}}$ instead of 60 ${M}_{\mathrm{Jup}}$), making many nondetections of lithium seemingly discrepant with model predictions. (Note that we use Cond model ages for plotting here because they are the only models that cover the full range of luminosities in our sample, and they tend to agree very well with the results from SM08 and BHAC models. We use the ages from our total-mass analysis, except for LHS 2397aAB, where we use the age from the individual-mass analysis of the secondary. For clarity, the unresolved multiples 2MASS J0920+3517B and 2MASS J0700+3157B are not shown. The two young systems that lack lithium information are Gl 569Bab and HD 130948BC.)

    Standard image High-resolution image

    Also shown in Figure 9 are the lithium-depletion mass limits predicted by BHAC models and by the Tucson models (Burrows et al. 1997). Although we do not use the Tucson models in our other model analysis, mainly because in addition to their non-gray atmosphere models they used gray atmospheres at the higher temperatures (including some temperatures relevant for our sample) in order to cover the full range of evolution, we include them here because they are the only other available evolutionary tracks that report lithium depletion. As expected, lithium nondetections in the most massive objects, LHS 2397aA (M8, 93 ± 4 ${M}_{\mathrm{Jup}}$) and the young (${271}_{-29}^{+22}$ Myr, BHAC) components of LP 349-25AB (M7+M8, 85 ± 4 ${M}_{\mathrm{Jup}}$ and 80 ± 3 ${M}_{\mathrm{Jup}}$), are all consistent with model-predicted lithium depletion. Likewise, the two systems with strong lithium detections comprise some of the lowest mass components in the sample (≲50 ${M}_{\mathrm{Jup}}$), so they are consistent with both sets of models that predict they should be abundant in lithium.

    The lowest mass objects with lithium not detected are the 59 ± 5 ${M}_{\mathrm{Jup}}$ and 41 ± 4 ${M}_{\mathrm{Jup}}$ components of DENIS J2252−1730AB (L4+T3.5; ${1.10}_{-0.18}^{+0.15}$ Gyr). Within the 1σ uncertainties, the primary is consistent with being strongly depleted according to the BHAC models, $\mathrm{log}(\mathrm{Li}/{\mathrm{Li}}_{\mathrm{init}})=-{2.1}_{-1.5}^{+0.7}$ dex (Cond), and the late-type secondary is likely too cool to possess monatomic lithium. In contrast, the Tucson models predict that DENIS J2252−1730A should have retained most of its lithium, $\mathrm{log}(\mathrm{Li}/{\mathrm{Li}}_{\mathrm{init}})=-{0.04}_{-0.06}^{+0.04}$ dex, and are only consistent with full depletion at 2.0σ. Five other objects in four systems with masses below 70 ${M}_{\mathrm{Jup}}$ (2MASS J2132+1341AB, 2MASS J0700+3157A, LHS 2397aB, 2MASS J1728+3948B) are all consistent at ≤1σ with full lithium depletion according to the BHAC models. In contrast, Tucson models predict that all of these objects could possess a substantial fraction (up to 100%) of their initial lithium within the 1σ errors. This can be seen visually in Figure 9 as the cluster of lithium nondetections around 60–70 ${M}_{\mathrm{Jup}}$ and 2 Gyr that mostly lie above the BHAC 99.9% depletion curve but that lie entirely below the Tucson 99.9% depletion curve.

    We have noted this tension between the predictions of the Lyon and Tucson models in our previous work on HD 130948BC (Dupuy et al. 2009b). Our large sample of individual masses below the substellar boundary suggests that lithium is destroyed via fusion at lower masses than the Tucson models predict but at masses consistent with the BHAC models. The fact that the Tucson models predict a higher mass limit for lithium fusion indicates that their central temperatures are lower, which can be understood as a consequence of their atmospheres being less opaque as discussed in detail in Section 3.2 of SM08. According to the BHAC models, at field ages of ≳1 Gyr, the lithium depletion boundary is at a mass of 60 ${M}_{\mathrm{Jup}}$. Some of the most interesting systems for testing lithium depletion are the young (400–800 Myr) binaries Gl 569Bab, HD 130948BC, and Gl 417BC, which span masses of ≈50–70 ${M}_{\mathrm{Jup}}$ but do not yet have directly measured individual masses and/or optical spectra. Individual masses are attainable in the future with absolute astrometry, either relative to their stellar hosts (Gl 569Bab, HD 130948BC) or from wide-field HST imaging (Gl 417BC). In addition, resolved optical spectra from HST/STIS would benefit the entire sample, enabling stricter tests of models than are currently possible with integrated-light spectra.

    7.3. Coevality Tests and the Mass–Luminosity Relation in the L/T Transition

    In principle, every binary in our sample can be thought of as a "mini cluster" of coeval, co-compositional objects that can be used to test model isochrones (e.g., Liu et al. 2010). The most straightforward approach is to frame this as a test of coevality. Given our directly measured individual masses and luminosities, we derive from models an age posterior distribution for each component (t1 and t2). We then examine the posterior distribution of the difference in ages, and the extent to which it is different from zero corresponds to a discrepancy in the mass–luminosity relation predicted by model isochrones. Because we use a Monte Carlo approach, covariances due to parameters in common between components (distance and total mass) are naturally accounted for in our analysis.

    Figure 10 displays the median and 1σ intervals of the posterior distributions for ${\rm{\Delta }}\mathrm{log}(t)\equiv \mathrm{log}{t}_{2}-\mathrm{log}{t}_{1}$. To examine how this coevality criterion depends on the underlying physical properties, we plot these values as a function of the secondary component mass (i.e., the component that is less luminous). At the highest masses, where the binaries are composed of two main-sequence stars that each have unconstrained age posteriors, the resulting coevality values are all ${\rm{\Delta }}\mathrm{log}t\approx 0.0\pm 0.5$ dex at 1σ. In other words, both stars are consistent with the full range of main-sequence ages (a few hundred Myr, depending on mass, to 10 Gyr, the maximum age of the models).

    Figure 10.

    Figure 10. Coevality determinations plotted as a function of the mass of the secondary component. For each system with directly measured individual masses and luminosities, we derive the age of each component from the models (t1 and t2). The data points here show the median and 1σ intervals in the posterior distribution of the logarithm of the difference in age, computed as $\mathrm{log}{t}_{2}-\mathrm{log}{t}_{1}$. Gray symbols indicate two of the likely unresolved multiples LP 415-20 and 2MASS J0700+3157. Binaries composed of main-sequence stars (secondary masses >70 ${M}_{\mathrm{Jup}}$ here) have essentially unconstrained ages and thus ${\rm{\Delta }}\mathrm{log}t\approx 0.0\pm 0.5$ dex (i.e., both stars consistent with any main sequence age at 1σ). Aside from the unresolved multiple 2MASS J0700+3157, all but two systems give consistent ages for the primary and secondary at ≲1σ. These are two binaries spanning the L/T transition that have the lowest mass secondaries. SDSS J0423−0414AB and SDSS J1052+4422AB are 3.0σ and 1.5σ, respectively, discrepant with coevality in the Cond models. In contrast, SM08 models that predict a shallower mass–luminosity relation in the L/T transition yield coeval ages for both systems.

    Standard image High-resolution image

    The more interesting coevality tests are for binaries composed of two brown dwarfs. All but two of these systems give consistent ages for the primary and secondary at ≲1σ across all models. The exceptions are two binaries spanning the L/T transition that have the lowest-mass secondaries of any objects in our sample. According to the Lyon Cond models, SDSS J0423−0414AB and SDSS J1052+4422AB are 3.0σ and 1.5σ discrepant with coevality, respectively. In contrast, SM08 models predict a shallower mass–luminosity relation in the L/T transition and thus yield coeval ages for both systems (consistent at 0.5σ and 0.2σ, respectively). The key difference between the Lyon models (Cond and Dusty) and the SM08 hybrid models is that SM08 prescribes an ad hoc change from a cloudy to a cloudless photosphere as the temperature drops from 1400 K to 1200 K. This results in luminosity dropping less quickly during and immediately following the transition from cloudy to cloudless atmospheres.

    We originally discussed this discrepancy between the SM08 and Lyon models for SDSS J1052+4422AB in Dupuy et al. (2015b). The case of SDSS J0423−0414AB (L6.5 ± 1.5 and T2.0 ± 0.5) is very similar to SDSS J1052+4422AB (L6.5 ± 1.5 and T1.5 ± 1.0). The primary masses are the same within the errors (${51.6}_{-2.5}^{+2.3}$ ${M}_{\mathrm{Jup}}$ and 51 ± 3 ${M}_{\mathrm{Jup}}$, respectively), but SDSS J0423−0414AB has a lower mass ratio (0.62 ± 0.04 compared to 0.78 ± 0.07) and luminosity ratio (${\rm{\Delta }}\mathrm{log}({L}_{\mathrm{bol}})={0.31}_{-0.08}^{+0.09}$ dex compared to 0.13 ± 0.08 dex). Thus, SDSS J0423−0414AB has even more leverage in our coevality test, and indeed it confirms our previous findings at even higher significance. The reason for the coevality test failure in both cases is that the secondary is more luminous (or equivalently the primary is less luminous) than predicted for their masses at a single age in Lyon models. According to the SM08 hybrid models, luminosity does not fade as quickly during and immediately following the transition from cloudy to cloud-free atmospheres. So according to the SM08 models, it is the secondary that is more luminous than expected (rather than a suppressed luminosity of the primary) because the secondary has already cooled through the L/T transition and lost its clouds.

    A new physical explanation for the L/T transition has been proposed by Tremblin et al. (2016), where the primary driver is a thermo-chemical instability in the carbon chemistry (CO/CH4). Our measurement of the mass–luminosity relation through the L/T transition will provide a key test of this new idea after it is implemented in evolutionary models in the future.

    7.4. Discovery of Triple Systems

    Our individual mass and luminosity measurements have revealed that some objects in our sample are likely triple systems where the higher-order multiplicity is unresolved by current observations. The most extreme case is 2MASS J0920+3517AB (L5.5+L9), where the total mass of 187 ± 11 ${M}_{\mathrm{Jup}}$ is high enough that it suggests the presence of an unresolved massive component. Our individual masses reveal that indeed the less luminous (by 0.06 ± 0.04 dex) component 2MASSJ0920+3517B is much more massive (${116}_{-8}^{+7}$ ${M}_{\mathrm{Jup}}$) than the more luminous 2MASS J0920+3517A (71 ± 5 ${M}_{\mathrm{Jup}}$). In fact, the anomalous mass of 2MASS J0920+3517B would be obvious on its own, given that its spectral type is L9.0 ± 1.5, and that models cannot reproduce its luminosity at such a high mass (Figure 11). 2MASS J0920+3517B can plausibly be explained by a simplistic model in which it is composed of two equal-luminosity components with equal masses of ${58}_{-4}^{+3}$ ${M}_{\mathrm{Jup}}$. The unresolved pair must be rather tight, as we have never observed an elongated PSF for the secondary relative to the primary, and the astrometric residuals about the resolved orbit fit have an rms of 0.5 mas for the better half (13 out of 25) of our measurements. This implies an inner orbit that is much smaller than the observed outer orbit (a = 68.15 ± 0.23 mas, 2.11 ± 0.04 au). We discuss this system in more detail in Appendix B.7.

    Figure 11.

    Figure 11. 

    Individual mass and luminosity measurements shown alongside models. These are all likely triples given that the data points significantly deviate from coevality (i.e., falling on a single isochrone). In the cases of 2MASS J0700+3157AB and 2MASS J0920+3517AB, the less luminous secondary is actually more massive than the more luminous primary, implying that the secondaries are in fact unresolved binaries. In the case of LP 415-20AB, the primary is much less luminous than expected given its mass, implying that it is an unresolved pair of lower luminosity, lower mass objects. (An extended version of this figure is available.)

    Standard image High-resolution image

      Unlike 2MASS J0920+3517AB, the total mass of the L3+L6.5 binary 2MASS J0700+3157AB (${141}_{-5}^{+4}$ ${M}_{\mathrm{Jup}}$) is not unusually high. Indeed, our total-mass model analysis readily finds a self-consistent way to apportion the mass of the two components (${76.8}_{-1.3}^{+1.5}$ ${M}_{\mathrm{Jup}}$ and ${66}_{-3}^{+4}$ ${M}_{\mathrm{Jup}}$, Cond) according to their quite different luminosities (${\rm{\Delta }}\mathrm{log}({L}_{\mathrm{bol}}/{L}_{\odot })\,=0.50\,\pm 0.06$ dex) at an age of ${2.1}_{-0.6}^{+0.4}$ Gyr. However, our directly measured masses are 68.0 ± 2.6 ${M}_{\mathrm{Jup}}$ and ${73.3}_{-3.0}^{+2.9}$ ${M}_{\mathrm{Jup}}$. The less luminous component is again the more massive one, implying that 2MASS J0700+3157B is in fact an unresolved binary. The model-derived ages for 2MASS J0700+3157A are ${0.76}_{-0.14}^{+0.09}$ Gyr (SM08) and ${1.01}_{-0.19}^{+0.13}$ Gyr (Cond). If we assume that 2MASS J0700+3157B is composed of equal-luminosity components with equal masses of ${36.7}_{-1.5}^{+1.4}$ ${M}_{\mathrm{Jup}}$, their model-derived ages are somewhat inconsistent with the primary's (${1.15}_{-0.12}^{+0.11}$ Gyr and ${0.83}_{-0.09}^{+0.08}$ Gyr, respectively), suggesting that this simplistic scenario is not likely. As with 2MASS J0920+3517AB, we have never observed evidence for PSF elongation of 2MASS J0700+3157B in our Keck LGS AO imaging, and the rms about the relative orbit fit is 0.22 mas for the better half (11 out of 22) of our measurements. Thus, the inner orbit in this system is likely very small relative to the outer orbit ($a={377}_{-6}^{+5}$ mas, 4.25 ± 0.08 au). We discuss this system in more detail in Appendix B.4.

      We have identified a third candidate triple system. The total mass of LP 415-20AB (${248}_{-29}^{+26}$ ${M}_{\mathrm{Jup}}$) is very high for such late spectral type components (M6+M8). Our individual masses show that this is because the primary is very massive (${156}_{-18}^{+17}$ ${M}_{\mathrm{Jup}}$), while the secondary's mass (${92}_{-18}^{+16}$ ${M}_{\mathrm{Jup}}$) is consistent with its luminosity according to the BHAC models. Given the relatively large individual mass uncertainties, it is unclear whether the putative unresolved component of LP 415-20A is another low-mass star or if it is a brown dwarf. It is less likely, but still possible, that the apparent discrepancy here is due to the rather large observational uncertainties. An improved parallax for this system, which is also notable as a possible member of the Hyades, will help clarify the situation as we discuss in more detail in Appendix B.2.

      2MASS J0700+3157 and 2MASS J0920+3517 bring the tally of high-order multiples composed entirely of likely brown dwarfs to five, and these are the first such triple systems with directly measured masses, semimajor axes, and eccentricities. The first triple brown dwarf identified was DENIS J020529.0−115925 (Bouy et al. 2005), where the primary (L5) is orbited by an L/T transition binary (L8+T0). Radigan et al. (2013) discovered the triple T dwarf system 2MASS J08381155+1511155, where the brightest component (T3) is orbited by two slightly fainter components (T3+T4.5). Stone et al. (2016) discovered that the primary in the young system VHS J125601.92−125723.9 (Gauza et al. 2015) is actually a binary, making this the only one of these hierarchical triples in which the inner pair contains the most massive components. (Note that the distance of VHS J1256−1257 is not secure, and the more massive inner pair could be composed of low-mass stars and not brown dwarfs.) The hierarchical mass ordering of the DENIS J0205−1159 and 2MASS J0838+1511 systems appear to be similar to 2MASS J0700+3157 and 2MASSJ0920+3517, but the relative sizes of the inner and outer orbits may be different. The inner and outer pairs of DENIS J0205−1159 have projected separations of 1.3 au and 7 au, respectively, and 2MASS J0838+1511 has projected separations of 2.5 au and 27 au. If the inner pairs of 2MASS J0700+3157 and 2MASS J0920+3517 were only a factor of 7 × –10× smaller than the orbit, then we would have either resolved them directly or possibly seen perturbations in our relative astrometry. Thus, it is likely that the semimajor axis ratios of 2MASS J0700+3157 and 2MASS J0920+3517 are much smaller than for DENIS J0205−1159 and 2MASS J0838+1511. Finally, we note that the eccentricity of the outer orbit is very low for 2MASS J0700+3157 (${0.017}_{-0.007}^{+0.005}$) and is also rather low for 2MASS J0920+3517 (${0.180}_{-0.007}^{+0.006}$). This would seem to rule out a violent dynamical origin for these triple systems.

      7.5. Effective Temperature Relations

      Without directly measured radii for field brown dwarfs of known spectral type, all previous works on spectral type–${T}_{\mathrm{eff}}$ relations have relied on assumptions about age combined with model radii (e.g., Golimowski et al. 2004a; Vrba et al. 2004; Stephens et al. 2009; Filippazzo et al. 2015). Our sample of visual binaries with directly measured luminosities and precise model-derived ages and radii allows us to examine spectral type–${T}_{\mathrm{eff}}$ relations without uncertainties in radius, due to the unknown ages or masses. We consider the effective temperatures derived from the BHAC and Cond models as a single "Lyon" relation, using BHAC results when available. We used results from our individual-mass analysis when possible, relying on our total-mass analysis for the three systems without astrometric mass ratios (Gl 417BC, Gl 569Bab, and HD 130948BC). We excluded the three likely unresolved binaries LP 415-20A, 2MASS J0700+3157B, and 2MASS J0920+3517B. LHS 2397aB is also excluded because it lacks a spectral type determination. This results in a sample of 40 objects with Lyon temperatures and 22 objects with SM08 temperatures. (SM08 results are only for the coolest objects; our highest SM08 model-derived ${T}_{\mathrm{eff}}$ is 2090 ± 50 K for 2MASS J1017+1308A.)

      Figure 12 shows our model-derived effective temperatures as a function of spectral type. We calculated second-order polynomial fits of ${T}_{\mathrm{eff}}$ as a function of spectral type, weighting by the quadrature average of the +1σ and −1σ uncertainties in ${T}_{\mathrm{eff}}$. The coefficients of these fits are given in Table 14. The Lyon fit covers spectral types of M7–T5 and ${T}_{\mathrm{eff}}\approx 1100\mbox{--}2800$ K, and the residuals have an rms scatter of 90 K. The SM08 fit covers spectral types of L1.5–T5 and ${T}_{\mathrm{eff}}\approx 1100\mbox{--}2100$ K, and the residuals have an rms scatter of 80 K. The two fits agree reasonably well in the overlapping spectral type range, with the main difference being that the SM08 fit gives ≈100 K cooler temperatures for L4–L7 dwarfs. This seems to be a reflection of the fact that the SM08 model-derived temperatures for these objects are indeed systematically lower, due to SM08 radii being ≈10% larger than Cond radii at these temperatures, and not simply a quirk of the polynomial fit. This is expected from the fact that the SM08 models have clouds while Cond models do not (e.g., see the discussion of cloudy versus clear models by Burrows et al. 2011). There are no significant outliers in the SM08 fit, and in the Lyon fit only Gl 569Bb appears to be ≳1.5σ discrepant. It has a model-derived temperature that is 310 K cooler than expected given its M9 spectral type and more consistent with L0.5–L1 types. Gl 569Ba is also slightly discrepant for its M8.5 spectral type (140 K cooler than the fit). Therefore, Gl 569Bb may appear discrepant because the system as a whole is a slight outlier in ${T}_{\mathrm{eff}}$ due to the shared distance uncertainty in addition to the spectral type of Gl 569Bb perhaps being slightly underestimated.

      Figure 12.

      Figure 12. Effective temperatures derived from the SM08models (top) and from the Lyon models (bottom; BHAC and Cond). Solid black lines show our second-order polynomials to the relations, having rms scatter about the fit of 90 K for the 40 objects with Lyon temperatures and 80 K for the 22 objects with SM08 temperatures. Gray lines indicate previous literature polynomial relations from Golimowski et al. (2004a, dotted) and Filippazzo et al. (2015, dashed).

      Standard image High-resolution image

      Table 14.  Coefficients of Polynomial Fits

      y x c0 c1 c2 c3 c4 c5 rms Valid x Range
      ${T}_{\mathrm{eff}}$ (Lyon) SpT 4251.0 −238.03 4.582 90 K M7–T5
      ${T}_{\mathrm{eff}}$ (SM08) SpT 4544.3 −284.52 6.001 80 K L1.5–T5
      MKO Photometric System
      $\mathrm{log}({L}_{\mathrm{bol}}/{L}_{\odot })$ MK −172.188 68.11147 −10.671188 0.8162709 −0.03068824 0.000454547 0.05 dexa 9.1–17.8 mag
      $\mathrm{log}({L}_{\mathrm{bol}}/{L}_{\odot })$ MH −13.282 3.46876 −0.351721 0.0106200 0.023 dex 9.6–13.3 mag
      ${J}_{\mathrm{MKO}}-{J}_{2\mathrm{MASS}}$ MK −2.284 0.55165 −0.043190 0.0010379 0.015 mag 9.1–17.0 mag
      ${H}_{\mathrm{MKO}}-{H}_{2\mathrm{MASS}}$ MH −0.051 0.00850 0.003 mag 9.6–13.4 mag
      ${H}_{\mathrm{MKO}}-{H}_{2\mathrm{MASS}}$ MK −0.702 0.15644 −0.010605 0.0002365 0.004 mag 9.1–17.0 mag
      ${K}_{\mathrm{MKO}}-{K}_{S,2\mathrm{MASS}}$ MK −0.048 0.00221 0.006 mag 9.1–13.0 mag
      ${K}_{\mathrm{MKO}}-{K}_{S,2\mathrm{MASS}}$ MK −1.526 0.11594 0.007 mag 13.0–14.2 mag
      ${K}_{\mathrm{MKO}}-{K}_{S,2\mathrm{MASS}}$ MK 0.054 0.00447 0.007 mag 14.2–17.0 mag
      ${K}_{{\rm{H}}2}$K MK −0.207 0.07241 −0.004693 0.025 mag 9.1–13.1 mag
      2MASS Photometric System
      $\mathrm{log}({L}_{\mathrm{bol}}/{L}_{\odot })$ ${M}_{{K}_{S}}$ −59.877 22.58776 −3.364101 0.2357643 −0.00785837 0.000098821 0.05 dexa 8.8–16.6 mag
      $\mathrm{log}({L}_{\mathrm{bol}}/{L}_{\odot })$ MH −10.426 2.74259 −0.290797 0.0089222 0.023 dex 9.2–13.3 mag
      ${J}_{\mathrm{MKO}}-{J}_{2\mathrm{MASS}}$ ${M}_{{K}_{S}}$ −2.747 0.66217 −0.051739 0.0012518 0.017 mag 9.1–16.6 mag
      ${H}_{\mathrm{MKO}}-{H}_{2\mathrm{MASS}}$ ${M}_{{K}_{S}}$ −0.754 0.16903 −0.011606 0.0002626 0.004 mag 9.1–16.6 mag
      ${H}_{\mathrm{MKO}}-{H}_{2\mathrm{MASS}}$ MH −0.050 0.00845 0.003 mag 9.6–13.4 mag
      ${K}_{\mathrm{MKO}}-{K}_{S,2\mathrm{MASS}}$ ${M}_{{K}_{S}}$ −0.048 0.00219 0.006 mag 9.1–12.9 mag
      ${K}_{\mathrm{MKO}}-{K}_{S,2\mathrm{MASS}}$ ${M}_{{K}_{S}}$ −1.413 0.10801 0.026 mag 12.9–14.2 mag
      ${K}_{\mathrm{MKO}}-{K}_{S,2\mathrm{MASS}}$ ${M}_{{K}_{S}}$ 0.088 0.00240 0.007 mag 14.2–17.0 mag
      ${K}_{{\rm{H}}2}$${K}_{S}$ ${M}_{{K}_{S}}$ −0.036 0.03407 −0.002824 0.021 mag 9.1–13.1 mag

      Notes. The coefficients are defined as

      where y and x are the quantities listed in the first two columns. For spectral types, M7 corresponds to 7.0 and T5 corresponds to 25.0.

      aThe luminosity relations have significantly different scatter at bright and faint magnitudes. For both, the scatter is 0.04 dex at ≤13.0 mag and 0.07 dex at >13.0 mag.

      Download table as:  ASCIITypeset image

      Figure 12 also shows literature polynomial relations from Golimowski et al. (2004a) and Filippazzo et al. (2015). Our mass-calibrated relation broadly agrees with this past work, except for the temperature of the L/T transition from Golimowski et al. (2004a), which is systematically high, likely due to the presence of unrecognized binaries in that early work (e.g., Burgasser et al. 2005; Liu et al. 2006). The relation of Filippazzo et al. (2015) is systematically lower than our Lyon relation by ≈90 K, which appears to be due to the fact that they used SM08 model radii. Our SM08 relation agrees very well with Filippazzo et al. (2015), except at the extreme late-type end where we find a 150 K warmer ${T}_{\mathrm{eff}}$ for spectral type T5. This seems to be due to their polynomial not capturing the near-plateau in temperatures through the L/T transition which in our data extends to at least a type of T5. Although our sample does not exactly span the same range in spectral type, the scatter about our polynomial relation (90 K for M7–T5) is somewhat smaller than previous relations spanning M6–T8 (124 K for Golimowski et al. 2004a; 113 K for Filippazzo et al. 2015).

      Interestingly, all six T dwarfs in our sample, ranging from T1.5–T5, have effective temperatures that are consistent with each other within the uncertainties, i.e., p(χ2) ≈ 0.5. The weighted average and rms of the model-derived temperatures are 1200 ± 60 K for Lyon and 1190 ± 60 K for SM08. The batch of objects at slightly earlier types are five L6.5–L8.5 dwarfs that also have internally consistent temperatures of 1475 ± 30 K (Lyon) and 1400 ± 25 K (SM08). In comparison, the only object in our sample without a spectral type determination is LHS 2397aB, which has model-derived temperatures of 1520 ± 40 K (Lyon) and 1440 ± 40 K (SM08), consistent with the warmer L6.5–L8.5 dwarfs. This is the first significant sample of objects with individually measured masses spanning the L/T transition. Their model-derived radii combined with our luminosities imply that the spectral type range of L6.5–T5 corresponds to a temperature range of only 200–300 K, with Lyon models favoring the slightly larger temperature range. The narrower ${T}_{\mathrm{eff}}$ range derived from the SM08 models is likely due to the fact that ${L}_{\mathrm{bol}}$ does not drop as steeply through the L/T transition, as discussed in the previous subsection.

      Figure 13 shows our sample on the CMD with the data points colored according to their model-derived temperatures. As expected, temperature correlates strongly with the location along the CMD sequence, consistent with temperature being a primary driver of the spectral energy distributions of ultracool dwarfs. Of course, our derived ${T}_{\mathrm{eff}}$ is directly correlated with the plotted absolute magnitudes that are also used to derive ${L}_{\mathrm{bol}}$. However, in the L/T transition, objects have similar absolute magnitudes and yet they span a relatively wide range of temperatures (≈1500 K to ≈1100 K), with objects bluer in JK having systematically cooler model-derived temperatures.

      Figure 13.

      Figure 13. Color–magnitude diagram (CMD) showing our dynamical mass sample with symbols colored according to their Lyon model-derived effective temperatures (BHAC when possible, otherwise Cond). Upward-pointing triangles indicate primary components, downward-pointing triangles indicate secondary components, and the field sequence is shown for reference as gray squares using the latest compilation of ultracool dwarf parallaxes at http://www.as.utexas.edu/~tdupuy/plx. As expected, temperature correlates strongly with the location along the CMD sequence, consistent with temperature being a primary driver of the spectral energy distributions of ultracool dwarfs. Objects at the faintest magnitudes (MJ = 14–15 mag) span temperatures of ≈1500 K to ≈1100 K, where objects with bluer JK colors have cooler model-derived temperatures.

      Standard image High-resolution image

      7.6. Age Distribution

      The age distribution of the field population provides information about the star formation history of the galaxy and is a key component in studies of substellar mass function in the solar neighborhood (e.g., Burningham et al. 2010; Kirkpatrick et al. 2012). Without directly measured fundamental properties, the age distribution of ultracool dwarfs has previously only been constrained in a statistical sense from kinematic or population synthesis studies (e.g., Burgasser 2004; Allen et al. 2005; Zapatero Osorio et al. 2007; Faherty et al. 2009). However, precise age information is potentially available for brown dwarfs because, unlike stars, they are continuously changing their most easily observable properties, namely luminosity and temperature. The most precise age-dating method within current capabilities is to combine mass and luminosity, two properties that can be measured very accurately as we have shown here and in our past work, and infer an age from evolutionary models (for a review of various brown dwarf age-dating techniques, see Burgasser 2009). Note that, as in all stellar astrophysics, age determinations for brown dwarfs are model dependent, so the accuracy of the derived ages will depend on how well evolutionary models predict substellar luminosity evolution. As we have found in our past work, this is not yet a solved problem, with potential issues at the factor of ≈2 level (e.g., Dupuy et al. 2009b, 2014, 2015b). However, given that our derived ages span nearly two orders of magnitude (∼250 Myr to ∼10 Gyr), such inaccuracies should not have significant influence on the broad trends in our sample.

      Figure 14 shows the distribution of system ages for the 10 binaries in our sample with at least one substellar component and thereby a well-determined age. When possible we use SM08 model-derived ages from our total-mass analysis; we only use our individual-mass analysis for the special cases of 2MASS J0700+3157AB, where the secondary is an unresolved binary, and LHS 2397aAB, where the primary is a main-sequence star. We also tested ages derived from Cond models and found that they tended to give slightly (≈10%) older ages. We exclude two systems from our analysis here because their presence in our sample is not independent of their age. HD 130948BC and Gl 569Bab were both discovered in targeted surveys of young stars, and so it would not be correct to include them in this sample of field objects that have different observational selection effects with respect to age. Unlike these two companion systems, Gl 417BC was originally identified on its own in 2MASS and was only later associated with the young star Gl 417 by Kirkpatrick et al. (2001). Likewise, the pre-main-sequence binary LP 349-25AB is not included here because it only has a precise age determination by virtue of its youth.

      Figure 14.

      Figure 14. Bottom: age distribution for the systems with well-determined ages (solid red line). The histogram is the mean of 104 posterior values for each system, resulting in non-integer values for each 1 Gyr-wide bin. The dotted line shows the age distribution of a synthesized population of brown dwarfs with masses of 30–70 ${M}_{\mathrm{Jup}}$ and assuming a constant star formation rate (but including the young-age skew due to dynamical heating that removes old objects from the Galactic midplane, according to the Besançon model of the solar neighborhood). Top: cumulative distribution functions computed for the median age of each system (thick red line) and for 100 randomly drawn posterior age values for each system (thin orange lines). The observed distribution is remarkably consistent with the simplistic synthesized population that assumes a constant star formation rate for the thin disk and a spectral type cut of T5, as is the case for our sample of brown dwarfs here.

      Standard image High-resolution image

      The median age of our sample of field brown dwarfs is 1.3 Gyr (2.3 Gyr mean), and the age interval containing 90% of the joint posterior distribution of all 10 systems is 0.4–4.2 Gyr. This age distribution is broadly consistent with or somewhat younger than previous statistical age distributions. From a kinematic analysis of 21 L and T dwarfs, Zapatero Osorio et al. (2007) found a statistical age of ${1.2}_{-0.7}^{+1.1}$ Gyr, in good agreement with our age distribution. (Their sample includes just one of our age-dated systems, 2MASS J1728+3948AB.) In contrast, Faherty et al. (2009) found an older statistical age range from a kinematic analysis of proper motions for a much larger sample of 184 L0–T8 dwarfs, ranging from 3–8 Gyr for the whole sample, or 2–4 Gyr excluding high tangential velocity (>100 km s−1) objects. (None of our systems have such high tangential velocities.) The Faherty et al. (2009) results are thus systematically older than, but marginally consistent with, our age distribution.

      Allen et al. (2005) found that brown dwarfs with spectral types ranging from L5 to early-T are expected to have a mean age of 3 Gyr, given a uniform prior on age from 0–10 Gyr and a nominal mass function of ${\rm{\Psi }}(m)\propto {m}^{-0.8}$. The fact that we find a younger mean age could imply that the solar neighborhood comprises objects with a slightly younger age distribution. To investigate this possible discrepancy we performed a population synthesis simulation using a power-law mass function that is consistent with recent works (${\rm{\Psi }}(m)\propto {m}^{0.5};$ Kirkpatrick et al. 2012; Burningham et al. 2013) for a range in mass of 30–70 ${M}_{\mathrm{Jup}}$, i.e., the lowest to highest masses covered by our age-dated sample. Rather than assume a constant input age distribution, we adopted the Besançon model for the solar neighborhood that assumes a constant star formation rate and accounts for Galactic dynamics in a self-consistent way (Robin et al. 2003).13 Because the Sun lies near the Galactic midplane, dynamical heating skews the age distribution toward younger ages as older stars are preferentially scattered away from the midplane over time. In our simulation, we assume that ages are distributed uniformly among each of the Besançon model age bins, where each bin contains a fraction of the total population as follows: 20.0% for 0.15–1 Gyr, 16.1% for 1–2 Gyr, 11.9% for 2–3 Gyr, 16.6% for 3–5 Gyr, 12.9% for 5–7 Gyr, and 14.4% for 7–10 Gyr. We drew random ages according to this distribution, restricting our population synthesis simulation to 0.3–10 Gyr in order to accurately represent our observed sample.

      After assigning masses and ages to each simulated object, we interpolated ${T}_{\mathrm{eff}}$ from the SM08 models, keeping only simulated objects that fall within the temperature range of our sample (1100–2100 K). As a check, the fractional breakdown in spectral types between ≤L6, L6.5–L8.5, and L9–T5 dwarfs was 0.22, 0.36, and 0.42, very similar to the actual breakdown of objects in our age-dated sample (0.28, 0.33, 0.39). Also note that by performing our simulation with the same evolutionary models used to derive ages for our sample, the two age distributions are self-consistent by construction. In other words, the simulation output is directly comparable to our age distribution, even if the models were to have large systematic uncertainties in the absolute ages.

      The age distribution resulting from our simple population synthesis simulation is shown in Figure 14. It has a median of 1.7 Gyr, a mean of 2.3 Gyr, and the interval 0.3–5.0 Gyr contains 90% of the distribution. The general shape of the distribution is quite similar to our observations, piling up at younger ages and tailing off at older ages. We checked if the input mass function might change this result, but using α = 0 or −1 instead of −0.5 only changed the final median age by ≈5%. The simulation predicts that we would have found ≈1 old system (>5 Gyr) in our sample of 10 binaries, and the fact that none of our sample is definitively that old is the main cause of the slight discrepancy between the population synthesis and our observations. Overall, however, our age distribution is remarkably consistent with our input assumption of a constant star formation rate.

      The simulated population accounts for the main selection effects in our sample, namely a limited spectral type range and the fact that we can only age-date brown dwarfs and not stars (i.e., objects with mass ≲70 ${M}_{\mathrm{Jup}}$). The spectral type selection should bias our sample against old ages, and indeed the simulation indicates that we are the most complete at young ages. There is one remaining selection effect that is difficult to quantify, the fact that the youngest binaries at a given spectral type will have the lowest masses, and lower system masses correspond to longer orbital periods for a given semimajor axis range. Our dynamical mass sample relies on robust orbit determinations and thus suffers incompleteness at long periods (and thereby low masses and young ages) because there has been insufficient time to constrain the longest period orbits. (See Section 7.8 for a detailed discussion of this selection effect.) The fact that our sample is biased in this way against young ages, and yet we still find an age distribution skewed slightly younger than our simulation suggests that this selection effect has only a marginal influence.

      Overall, our population synthesis simulation demonstrates that our observed age distribution is consistent with a constant star formation rate in the Galactic disk. We find that the age distribution of M7–T5 dwarfs in the solar neighborhood is somewhat younger than in previous works, but we determine that this can be explained by accounting for dynamical heating in the disk that boosts the scale height of older stars, removing them from the local volume. As a test, we performed a second simulation with a uniform age input (i.e., constant star formation rate without dynamical heating), and this skewed the output median age older by almost a factor of two. In the future it will be possible to extend to older ages this empirical determination of the age distribution in the solar neighborhood with dynamical mass measurements for even cooler brown dwarfs (e.g., Dupuy et al. 2015a).

      7.7. Eccentricity Distribution

      We have previously examined the eccentricities of ultracool binaries and the implications for formation models in Dupuy & Liu (2011). The sample we used previously included eleven visual binaries, four spectroscopic binaries, and one eclipsing binary, and we had to contend with potential selection effects in this sample due to discovery bias (eccentric visual binaries are easier to discover with poor angular resolution), our own bias in selecting which binaries to monitor, and whether eccentricity might make some binaries more difficult to yield orbit determinations. Our new sample is not only much larger (10 new orbits), but we are also in a better position to address issues regarding observer and orbit-fitting biases. This is because we present here our entire monitoring sample with orbit fits for all binaries. The only selection is therefore based on the quality of the resulting fits, and this allows us to robustly quantify the completeness of our sample.

      As described in detail in Section 2, our input sample of 31 binaries is, to the best of our knowledge, complete for binaries with projected separations at discovery of ≤6 au, given our other non-orbit related selection criteria (spectral types M6.5–T5.5, d ≤ 40 pc, observable from Maunakea with Keck LGS AO). Figure 15 shows all orbit-fitting results for our sample, including binaries with marginally or poorly determined orbits, as well as visual binary orbits from the literature. As expected, the binaries with the widest separations at discovery have turned out to have longer orbital periods. Because we are hampered by the available observational time baseline for longer period binaries, many of these do not have well-determined orbits. Only one binary with P > 30 years has a well-determined orbit (2MASS J1728+3948AB); all the rest are marginal or poor. The only binary at P < 30 years that does not have a well-determined orbit is SDSS J0926+5847AB (e = 0.00–0.58 at 2σ). This is a pathological case where the orbit is very close to being viewed edge-on, and the phase coverage from our three-year HST program does not allow us to uniquely determine the orbit. It is likely that if SDSS J0926+5847AB were observable with Keck LGS AO, then we would have better phase coverage over a longer time baseline that would enable an orbit fit. Therefore, we conclude that for periods <30 years, our sample of orbit determinations is effectively complete, including 22 binaries from our sample and 3 binaries from the literature.

      Figure 15.

      Figure 15. Orbital period of our sample binaries plotted as a function of projected separation at discovery. Symbols are colored according to the quality of the orbit determination: good (black), marginal (orange), and poor (red). Open gray symbols indicate results from the literature for LHS 1070BC, $\epsilon $ Ind Bab, and LSPM J1314+1320AB. The quality of our orbit determinations are almost entirely determined by having sufficient observational time coverage, so the longest-period orbits tend to be poorly determined. The only P < 30 year binary with a poor orbit is 2MASS J0926+5847AB, which is seen nearly edge-on and lacks the phase coverage from our three-year HST program needed to robustly determine the orbit from such a challenging viewing angle. Therefore, we conclude that a cut of P < 30 years (dotted line) yields an effectively complete sample of orbit determinations to be used in a statistical study of orbital eccentricities. Most orbits have error bars smaller than the plotting symbols.

      Standard image High-resolution image

      A complete sample can still be biased, with the key effect here being "discovery bias," which is explored in detail in Section 3.1.1 of Dupuy & Liu (2011). Briefly, every imaging survey has an inner working angle (IWA), inside of which binary companions are not detectable. Given a binary with semimajor axis a, the projected separation will sometimes be inside of this limit, depending on the inclination, eccentricity, and epoch of observation. When IWA ≪ a, the probability of detecting any binary approaches unity. When IWA > a, circular binaries are completely undetectable while more eccentric binaries are more favorable to detect because the apoastron distance is $a\times (1+e)$ (see Figure 1 of Dupuy & Liu 2011). By excluding binaries discovered very near the IWA of a survey, we can minimize the impact of this discovery bias. Figure 16 shows the ratio of IWA/a for binaries in our sample. The highest IWA/a ratios are 1.58 and 1.24 for 2MASS J1047+4026AB and LP 415-20AB, respectively, and indeed both of these binaries are quite eccentric (0.7485 ± 0.0013 and ${0.706}_{-0.012}^{+0.011}$). The next highest are IWA/a = 0.88 for 2MASS J0920+3517AB ($e={0.180}_{-0.007}^{+0.006}$) and IWA/a = 0.69 for LP 349-25AB ($e={0.0468}_{-0.0018}^{+0.0019}$). Based on the simulations of Dupuy & Liu (2011), we adopt a cutoff of IWA/a < 0.75 to mitigate discovery bias in our sample, and this includes 2MASS J1047+4026AB, LP 415-20AB, and the more marginal case of 2MASS J0920+3517AB. Eccentricities for our entire sample, along with the three other published visual binary orbits, are shown in Figure 17 as a function of semimajor axis.

      Figure 16.

      Figure 16. Ratio of the inner working angle (IWA) of the discovery data for each binary to its semimajor axis (a). For a given a, a survey with IWA ≪ a is sensitive to binaries of all eccentricities, while a survey with $\mathrm{IWA}\gt a$ is strongly biased toward discovering eccentric binaries. According to simulations from Dupuy & Liu (2011), a cutoff of IWA/a < 0.75 will result in minimal bias in the resulting eccentricity distribution. Therefore, we exclude the binaries 2MASS J1047+4026AB, LP 415-20AB, and 2MASS J0920+3517AB (gray diamonds) from our statistical analysis. The first two of these have IWA/a > 1 and indeed turn out to have quite eccentric (≥0.7) orbits. Most orbits have error bars smaller than the plotting symbols.

      Standard image High-resolution image
      Figure 17.

      Figure 17. Eccentricity as a function of semimajor axis for all orbits in our sample. Symbol shapes indicate orbits that are simple binaries (circles), inner pairs of hierarchical triples (upward-pointing triangles), outer pairs of hierarchical triples (downward-pointing triangles), or an inner pair in a quadruple (square). Symbols are colored according to the quality of the orbit determination: good (black), marginal (orange), and poor (red). Open gray symbols indicate results from the literature for LHS 1070BC, $\epsilon $ Ind Bab, and LSPM J1314+1320AB. Objects expected to be impacted by discovery bias (IWA/a > 0.75) are enclosed by large gray diamonds. No significant trends in eccentricity with semimajor axis are apparent. Most orbits have error bars smaller than the plotting symbols.

      Standard image High-resolution image

      After making the cuts described above, the final complete de-biased sample comprises 22 ultracool visual binaries. In addition, we examine published orbits for various unresolved binaries: three astrometric orbits (Sahlmann et al. 2015a, 2015b; Koren et al. 2016), three spectroscopic orbits (Basri & Martín 1999; Joergens et al. 2010; Burgasser et al. 2016), and one double-lined eclipsing orbit for a very young brown dwarf binary in Orion (Stassun et al. 2006). Figure 18 shows all of these results, spanning more than three orders of magnitude in period. The unresolved binaries all have modest eccentricities (0.2–0.6), whereas our sample has a number of very eccentric (>0.7) and nearly circular (0.0–0.1) systems. The lack of low eccentricities (<0.2) in the seven unresolved binaries is the most striking, as nearly half (10 of 22) of the visual binaries have e < 0.2. More short-period binaries are needed to determine the significance of this apparent difference, given the current small sample with short periods.

      Figure 18.

      Figure 18. Eccentricity as a function of orbital period for binaries in our de-biased sample (filled red) and from the literature (open black). Symbol shapes indicate orbits that are simple binaries (circles), inner pairs of hierarchical triples (upward-pointing triangles), or outer pairs of hierarchical triples (downward-pointing triangles). Our sample and some published orbits have error bars smaller than the plotting symbols. Literature results for spectroscopic, astrometric, and eclipsing binaries tend to have short periods and modest eccentricities (0.2–0.6). In contrast, visual binaries at longer periods have some very eccentric orbits (≥0.7) and a substantial fraction of low-e orbits (10 of 22 have e = 0.0–0.2). Whether this is indicative of an actual difference in orbital properties will require a larger sample of short-period binaries.

      Standard image High-resolution image

      Figure 19 shows the eccentricity distribution of our de-biased visual binary sample. The abundance of low eccentricity orbits is the clearest result, reinforcing the conclusion of Dupuy & Liu (2011) that very low-mass binaries are very inconsistent with the high eccentricity orbits found in simulations from Stamatellos & Whitworth (2009) but consistent with the more modest eccentricities predicted from the simulations of Bate (2009). In fact, our larger sample here is less consistent with the Bate (2009) results that do not predict as many nearly circular orbits. For both sets of simulations, it is possible that the influence of gas over timescales longer than the simulated durations would damp eccentricities further and bring the simulations into better agreement with our observations.

      Figure 19.

      Figure 19. Bottom: eccentricity distribution of our de-biased visual binary sample (P < 30 years). Top: cumulative distribution functions computed for the median eccentricity of each orbit (thick red line) and for 100 randomly drawn posterior values for each orbit (thin orange lines). Almost half of the orbits (10 of 22) have low eccentricities (0.0–0.2).

      Standard image High-resolution image

      Compared to the extensive orbital information available for solar-type binaries spanning ≈6 orders of magnitude in period (e.g., Duquennoy & Mayor 1991; Raghavan et al. 2010), our visual binary sample covers a relatively narrow range of periods (≲1 order of magnitude). However, one striking feature of the recent large Raghavan et al. (2010) compilation of solar-type multiple systems (127 eccentricities) is that for a wide range of periods (∼102–106 days), there are no binary orbits with e < 0.1 and only three orbits in triple or quadruple systems with e < 0.1. In contrast, our de-biased visual binary sample has 5 out of 22 orbits with e < 0.1. Two of these are orbits in triple systems, where one is an "inner" orbit (i.e., the orbit of the tighter pair in the hierarchical triple; HD 130948BC) and one is an "outer" outer (2MASS J0700+3157AB). The other three all seem to be binaries with no other companions (LP 349-25AB, 2MASS J1534−2952AB, and 2MASS J2206−2047AB). This discrepancy between the samples of orbits of very low-mass and solar-type objects was first noted in Dupuy & Liu (2011), and it is even stronger now after roughly doubling the sample size of visual binaries. For example, the Fisher exact test comparing the number of binaries above and below e = 0.1 in our sample to Raghavan et al. (2010) gives a p-value of 0.0019, implying the two samples are inconsistent at 3.1σ. This difference in low-e orbits could be due to differences in initial conditions (i.e., binary formation is different at very low masses) and/or early evolution (i.e., different physical processes at work after formation). In the latter case, if dissipative gas disks are longer lived for very low-mass objects than solar-type stars, this could potentially result in a higher fraction of nearly circular orbits even if the initial conditions are the same for both samples.

      7.8. The Lowest Mass Brown Dwarfs in the Sample?

      As described in Section 4.1, we have excluded some binaries from our analysis because of their uncertain orbits. This is largely due to poor orbital coverage given their apparently long orbital periods relative to our observational time baseline. Since our sample was selected to have the smallest semimajor axes possible, some of these binaries are likely to be the lowest mass systems in our sample. Indeed, if we take our orbit determinations at face value, then 2MASS J0850+1057AB, SDSS J1021−0304AB, SDSS J1534+1615AB, and SDSSJ2052−1609AB all have lower total masses than the lowest mass object used in our analysis (SDSS J0423−0414AB, ${M}_{\mathrm{tot}}=83\pm 3$ ${M}_{\mathrm{Jup}}$).

      2MASS J0850+1057AB (L6.5+L8.5) has a total mass of only 54 ± 8 ${M}_{\mathrm{Jup}}$, and thus perhaps component masses of 20–30 ${M}_{\mathrm{Jup}}$. Given the component luminosities, this would correspond to an age of 150–400 Myr according to both the Lyon and SM08 models. The 2MASS J0850+1057 system has previously been suggested to be young based on having an M-dwarf companion (NLTT 20346; 4farcm13 or 8000 au away) with a young X-ray activity age (Faherty et al. 2011) and based on 2MASS J0850+1057A being bright at the J-band for its spectral type (Burgasser et al. 2011). However, the physical association of NLTT 20346 with the 2MASS J0850+1057 system is questionable (see Section 6.4 of Dupuy & Liu 2012), and the discovery and characterization of young late-L dwarfs indicate that they tend to be fainter at the J-band, not brighter (e.g., Faherty et al. 2012; Liu et al. 2013; Liu et al. 2016). Our improved proper motion measurement for the 2MASS J0850+1057 system is consistent with our previous findings, discrepant by 45 mas yr−1 (6.3σ) with the Faherty et al. (2011) proper motion for NLTT 20346, implying that the wide pair is not likely to be physically associated. Burgasser et al. (2011) also suggested that 2MASS J0850+1057A could be an unresolved binary based on its unusually bright J- and K-band absolute magnitudes given its spectral type of L7, making this a triple system of ultracool dwarfs. Our distance (31.8 ± 0.6 pc instead of 38 ± 6 pc) and spectral type of L6.5 ± 1.0 give absolute magnitudes comparable to other L5–L7 dwarfs (see Table 15 of Dupuy & Liu 2012), reducing the reason to suspect 2MASS J0850+1057A is an unresolved binary. Moreover, if our low total mass is accurate, then the component masses (if a triple system) would be 10–15 ${M}_{\mathrm{Jup}}$, making the system even younger. This seems unlikely given that the integrated-light spectrum lacks spectral similarities to other comparably young late-L dwarfs (e.g., Liu et al. 2013; Gizis et al. 2015).

      SDSS J1021−0304AB (T0+T5, ${M}_{\mathrm{tot}}={52}_{-7}^{+6}$ ${M}_{\mathrm{Jup}}$) and SDSS J1534+1615AB (T0+T5.5, ${M}_{\mathrm{tot}}={46}_{-7}^{+6}$ ${M}_{\mathrm{Jup}}$) have the most precise masses of any of the systems with poorly determined orbits. For plausible mass ratios, the implied component masses of these binaries would be 20–30 ${M}_{\mathrm{Jup}}$. With orbital period posteriors of ${86}_{-17}^{+13}$ years and ${58}_{-24}^{+39}$ years, respectively, it is unlikely that we will achieve significant improvement in their orbit determinations in the near future. SDSS J2052−1609AB (L8.5+T1.5, ${M}_{\mathrm{tot}}={69}_{-20}^{+14}$ ${M}_{\mathrm{Jup}}$) has a very uncertain total mass but the best chance of yielding a robust orbit in the near term, as its derived period is ${44}_{-14}^{+10}$ years and our observations already cover 20% of this.

      8. Conclusions

      We present astrometric monitoring of a well-defined sample of 31 ultracool dwarf visual binaries (component spectral types M7–T5) from our ground- and space-based observing programs spanning nearly a decade in time baseline. We determine robust orbits using resolved astrometry by combining our Keck laser guide star adaptive optics imaging and aperture-masking interferometry observations, HST imaging, and other published archival data. Our unresolved infrared imaging from CFHT/WIRCam allows us to measure precise parallactic distances and constrain the photocenter orbital motion of our binaries. We perform a simultaneous MCMC analysis of both resolved and unresolved astrometry in order to accurately determine posterior distributions of all 13 astrometric parameters. By careful consideration of our posteriors, we propose orbit quality metrics that separate robustly determined orbits (23 systems) from marginal and poorly determined orbits that typically have long orbital periods (≳50 years) and thus lack sufficient observational time baseline. Our combined orbit and parallax results directly yield dynamical total masses (median precision 5%, as good as 1.1%), and combining our measured infrared flux ratios with the photocenter orbit constraints enables determination of individual masses for 19 systems (median precision 7%, as good as 1.7%).

      We perform empirical tests of evolutionary models based solely on the observed properties of our sample. Also, we critically examine the fundamental properties of objects from the bottom of the main sequence through the L/T transition by deriving precise ages, radii, etc. from the models. For this analysis, we develop new relations for deriving bolometric luminosity from H- and K-band absolute magnitudes, as well as a method for inferring parameters from models given our mass and luminosity constraints. Unlike our past work that used a simple interpolation of models, our new approach is based on the Monte Carlo technique of rejection sampling and allows for explicit definitions of age priors as well as proper treatment of cases where an object of a given mass might have the same luminosity at multiple times during its evolution (e.g., a star on the main sequence or a deuterium-burning brown dwarf). We summarize our findings as follows.

      • 1.  
        We perform a novel empirical test of the mass limit for hydrogen fusion in stars by examining the maximum mass of the coolest objects in our sample. Among all late-L and T dwarfs, we find no objects more massive than 70 ${M}_{\mathrm{Jup}}$, implying a substellar boundary that is lower in mass than that found in some previous works (75 ${M}_{\mathrm{Jup}}$ has often been quoted), but reasonably consistent with the latest BHAC evolutionary tracks. Among objects more massive than 70 ${M}_{\mathrm{Jup}}$, the latest spectral types are L3–L5, implying that any object with a later spectral type than this is likely to be a brown dwarf independent of any model assumptions.
      • 2.  
        We examine the mass limit for lithium fusion in brown dwarfs. Our sample has 13 systems with at least one substellar component, and two of these have integrated-light Li i absorption reported in the literature while seven have published nondetections of lithium. After considering the potential for chemical depletion of monatomic lithium into molecular form (e.g., LiCl, LiOH), the observations are fully consistent with predictions of lithium depletion from Lyon evolutionary models (Baraffe et al. 2015). In contrast, Tucson models (Burrows et al. 1997) predict that lithium should be present at detectable levels in some of our more massive binary components (60–65 ${M}_{\mathrm{Jup}}$), but this is inconsistent with numerous Li i nondetections. The most massive objects in systems with lithium detected (SDSS J0423−0414A and Gl 417B) have masses of ≈52 ${M}_{\mathrm{Jup}}$, consistent with (in fact, much lower than) the BHAC model-predicted lithium depletion boundary of 60 ${M}_{\mathrm{Jup}}$ at ages ≳1 Gyr.
      • 3.  
        We perform coevality tests, deriving ages for each binary component independently and checking for consistency to test evolutionary model isochrones. All binaries are consistent with predictions from the various models we tested, except for two binaries spanning the L/T transition that have the lowest mass secondary components in our sample (SDSS J0423−0414AB and SDSS J1052+4422AB). For these two systems, our observations show a smaller difference in luminosity between the binary components, given their difference in mass, than predicted by the Lyon models (3.0σ and 1.5σ, respectively). However, both binaries are consistent with the SM08 hybrid models that assume a gradual loss of cloud opacity as objects cool from 1400 K to 1200 K, which has the side effect of making the mass–luminosity relation shallower through the L/T transition. This reinforces the results of Dupuy et al. (2015b).
      • 4.  
        We discover two new brown dwarf triple systems, where the inner pairs are identified via direct mass measurements but not spatially resolved. 2MASS J0700+3157B and 2MASS J0920+3517B are each the fainter component of a visual binary but are significantly more massive than the brighter component, implying that they are each unresolved, roughly equal-mass binaries. A third system, LP 415-20AB, has an unusually massive primary (${156}_{-18}^{+17}$ ${M}_{\mathrm{Jup}}$) given its M6 spectral type, making this a third candidate triple in our sample. 2MASS J0700+3157 and 2MASS J0920+3517 are only the fourth and fifth known triple-brown-dwarf systems known, and they are the only such systems with directly measured masses, semimajor axes, and eccentricities, information that could be used to understand their origins.
      • 5.  
        We construct a spectral type–effective temperature relation for ultracool dwarfs that is based on model radii free of assumptions about mass or age. We fit a second-order polynomial to temperatures derived from the Lyon models (a sample that spans M7–T5 and 1100–2800 K) as well as the SM08 models (valid over a narrower range of L1.5–T5 and 1100–2100 K). The scatter about the fit is 90 K for the Lyon models and 80 K for the SM08 models. Examining individual subsets of objects, we find that all six T1.5–T5 dwarfs have consistent temperatures within their uncertainties (weighted average and rms of 1190 ± 60 K using the SM08 models), and the five L6.5–L8.5 dwarfs likewise have internally consistent temperatures (1400 ± 25 K using the SM08 models).
      • 6.  
        The 10 systems in our sample with precisely determined ages (and provenances free of explicit age bias) allow us to directly measure the age distribution of brown dwarfs in the solar neighborhood for the first time. The median age of our sample is 1.3 Gyr, and the mean is 2.3 Gyr. Compared to previous statistical age constraints on the field population, our age distribution is either consistent (Zapatero Osorio et al. 2007) or systematically younger (Faherty et al. 2009), and it is also younger than previous modeling of the solar neighborhood (e.g., Allen et al. 2005). We performed a population synthesis simulation designed to mimic our sample in mass and spectral type range, using as input a constant star formation history. We find good agreement with our observations only after accounting for dynamical heating in the Galactic disk, which preferentially removes older stars from the solar neighborhood. Thus, our sample is consistent with a constant star formation rate in the Galactic disk, at least over the relatively young ages that we probe (≲4 Gyr).
      • 7.  
        We have revisited the eccentricity distribution of very low-mass binaries. In general, our results reinforce the conclusions of Dupuy & Liu (2011), now with a larger sample and more robustly quantified completeness. Perhaps the most striking feature of our visual binaries is the preponderance of low-eccentricity orbits. Out of 22 orbits in our de-biased sample, 10 have e < 0.2 and five have e < 0.1. This is in stark contrast to solar-type binaries and possibly also the relatively small sample of seven short-period, unresolved ultracool binaries with spectroscopic or astrometric orbits. We speculate that low-eccentricity orbits could reflect more extensive tidal damping during the early evolution of ultracool binaries if their gas disks are longer lived than more massive stars.

      Our individual masses provide a direct window into substellar astrophysics, empirically quantifying the mass limits for hydrogen (and lithium) fusion. The unique ability of binaries to serve as "mini clusters" has provided the strongest tests of the mass–luminosity relation in the poorly understood L/T transition. Indeed, a recent theoretical work by Tremblin et al. (2016) has called into question the primacy of clouds in driving changes in the most readily observable properties (colors, magnitudes, and spectral types) across the L/T transition. Our results highlight how the modeling of surface properties can have a significant influence on the bulk evolution of brown dwarfs, as the only available models that match our mass–luminosity relation data are those that treat the L/T transition as a change in cloud properties. Evolutionary models adopting the alternative hypothesis that the L/T transition is driven by a CO/CH4 thermo-chemical instability are needed to test this idea against our observations.

      Our sample lays the groundwork for future tests of models using objects of known mass. For example, objects from our sample will have a significant advantage for testing model atmospheres as they provide a joint constraint on ${T}_{\mathrm{eff}}$ and $\mathrm{log}(g)$ (via radius) given the directly measured masses and luminosities. To enable such tests, resolved spectroscopy of our binaries will need to be obtained, and this is within reach of existing and future high angular-resolution facilities. Moreover, resolved optical spectroscopy will be key in strengthening our tests of lithium depletion, as our current tests rely on integrated-light spectroscopy alone. In the further future, dynamical masses for late-T and Y dwarf binaries (e.g., Liu et al. 2011; Dupuy et al. 2015a) will open the door to new tests of models. Such measurements will reveal both the older analogs of the ≳30 ${M}_{\mathrm{Jup}}$ brown dwarfs in our current sample, as well as pushing to much lower masses, perhaps even to the mass range shared by directly imaged planets (5–15 ${M}_{\mathrm{Jup}}$).

      We are deeply grateful to Michael Ireland, especially for his early contributions to this project, which included training the author on analysis of masking data from NIRC2, as well as very helpful comments on this manuscript. This work would not have been possible without the excellent, long-running support of CFHT staff past and present, including Loïc Albert, Todd Burdullis, Pascal Fouqué Nadine Manset, Karun Thanjavur, Kanoa Withington, and all service observers who have executed our program over the last nine years. It is a pleasure to thank the Keck Observatory staff for assistance with the Keck LGS AO observing, including Joel Aycock, Randy Campbell, Al Conrad, Scott Dahm, Greg Doppmann, Heather Hershley, Marc Kassis, Jim Lyke, Jason McIlroy, Carolyn Jordan, Gary Punawai, Julie Riviera, Luca Rizzi, Terry Stickel, Hien Tran, and Cynthia Wilburn. We thank Isabelle Baraffe, Didier Saumon, and Mark Marley for providing expanded model grids and helpful discussions about substellar evolution; and the anonymous referee for a thoughtful and timely review. T.J.D. is indebted to many people for providing useful discussions over the course of this project, including Katelyn Allers, Josh Carter, Rio Jennings, Adam Kraus, Andrew Mann, and Aaron Rizzuto. We are also very grateful to Andrew Mann for assistance in validating our methods for determining bolometric luminosities. This work was supported by multiple NASA Keck PI Data Awards, administered by the NASA Exoplanet Science Institute. Support for our long-term HST program (GO-11593, GO-12317, and GO-12661) was provided by NASA through a grant from the Space Telescope Science Institute, which is operated by the Association of Universities for Research in Astronomy, Inc., under NASA contract NAS 5-26555. T.J.D. acknowledges support from NASA Exoplanet Research Program grant NNX15AD23G. M.C.L. acknowledges support from NSF grant AST09-09222 and AST15-18339. Our research has employed NASA's Astrophysical Data System; Posse East; the SIMBAD database operated at CDS, Strasbourg, France; the 2MASS data products; and the SpeX Prism Spectral Libraries, maintained by Adam Burgasser at http://www.browndwarfs.org/spexprism. Finally, the authors wish to recognize and acknowledge the very significant cultural role and reverence that the summit of Maunakea has always had within the indigenous Hawaiian community. We are most fortunate to have the opportunity to conduct observations from this mountain.

      Facilities: HST (WFC3) - , Keck:II (LGS AO - , NIRC2) - , CFHT (WIRCam) - , IRTF (SpeX). -

      Appendix A: Empirical Relations with Absolute Magnitude

      Although there are many useful empirical relations in the literature based on spectral type, many tight binary components do not have directly measured spectra. However, we do have accurate near-infrared absolute magnitudes for all of our binaries from parallactic distances and resolved photometry. We will soon have many more late-M and L dwarfs with distances from Gaia that will not immediately have spectra available. Therefore, we have derived empirical relations of various useful quantities with respect to absolute magnitude using the existing sample of ultracool dwarfs with parallaxes. Because J- and H-band absolute magnitudes do not change monotonically through the L/T transition (e.g., Dupuy & Liu 2012), we primarily derive these relations using K-band absolute magnitudes on the 2MASS and MKO photometric systems. We also examined some relations as a function of H-band absolute magnitude for MH < 13.3 mag objects for which the relations have lower scatter in luminosity and spectral type and the color changes monotonically.

      A.1. 2MASS-to-MKO Photometric Conversion

      In order to report complete photometry across both MKO and 2MASS photometric systems for our sample binaries, we compute photometric system conversions based solely on K-band absolute magnitude. We used our compilation of ultracool dwarf parallaxes available online,14 selecting only single objects with a published spectrum, a parallax error of <15%, and a "null" object flag, meaning that there is nothing unusual about the object. We use synthesized 2MASS/MKO photometric conversions from Dupuy & Liu (2012) for each of the 46 objects in each of the J-, H-, and K-bands. For the J- and H-bands, we found that a third-order polynomial provided a good fit to the data. The K-band conversion undergoes a steep jump between MK ≈ 13–14 mag that cannot be well-approximated by a single polynomial. Therefore, we computed a three-part piecewise linear fit to the K-band conversion as a function of MK. In addition to these relations as a function of MK, we also computed an H-band conversion as a function of MH. The polynomial coefficients are given in Table 14 and the fits are displayed in Figure 20. The typical rms about the fits ranges from 0.003 mag for the H-band to 0.015 mag for the J-band.

      Figure 20.

      Figure 20. Polynomial relations between MKO-to-2MASS photometric conversions and absolute magnitude. The K-band is a three-part piecewise linear fit, while others are single polynomial fits. The input data used to derive these fits (black points) are computed by synthetic photometry as described in Appendices A.1 and A.2. We use these relations to derive JHK photometry on both 2MASS and MKO systems in cases where it is not measured directly.

      Standard image High-resolution image

      A.2. Broadband K to Narrowband CFHT/WIRCam ${K}_{{\rm{H}}2}$ Conversion

      Determining mass ratios from photocenter motion requires knowing the flux ratios in the imaging bandpass. Our absolute astrometry for the brightest targets observed with CFHT/WIRCam was obtained in a narrow K-band filter, ${K}_{{\rm{H}}2}$, centered on 2.122 μm with a bandwidth of 0.032 μm. We do not have resolved photometry for these binaries in this specialized filter, so we must synthesize it from our broadband K photometry in order to determine the flux ratios of our binaries in our CFHT imaging.

      We used the same sample of objects selected from our online parallax compilation as described in the previous section. We used the filter curve provided by CFHT to synthesize ${K}_{{\rm{H}}2}$ photometry,15 and we assume that the difference between this scan done in the lab and the transmission at colder operational temperatures is negligible for our purposes. We computed ${K}_{{\rm{H}}2}$K and ${K}_{{\rm{H}}2}$${K}_{S}$ colors for these objects and found a smooth relation for objects with MK < 13.1 mag, corresponding roughly to spectral types <L9. (The smooth relations break down for later types as the ${K}_{{\rm{H}}2}$ filter samples part of the K-band flux peak of T dwarfs while their broadband K flux plummets.) We fit a second-order polynomial as a function of MK for 25 objects. The coefficients are given in Table 14 and the fits are displayed in Figure 20. The rms about these fits were 0.025 mag for K and 0.021 mag for ${K}_{S}$.

      A.3. Bolometric Luminosity

      Determining the component luminosities using our resolved photometry is a critical part of our model tests. We have used the large sample of luminosities from Filippazzo et al. (2015) to derive polynomial relations between luminosity and absolute magnitude. We used only objects with parallactic distances, nominal ages ≳0.5 Gyr, and normal properties (i.e., no subdwarfs, young objects, or objects dubbed peculiar in their spectral type). We supplemented the observed magnitudes reported by Filippazzo et al. (2015) by applying our 2MASS/MKO conversions from above. For any object with either a 2MASS or MKO measurement, we derived the complementary magnitude using the K-band absolute magnitude. For objects with both 2MASS and MKO measurements, we used the more precise magnitude to derive the complementary magnitude, supplanting the directly observed value. This results in a sample of 126 normal field objects with luminosities and K-band absolute magnitudes. (Since our 2MASS/MKO relations are only valid between MK = 9.1–17.0 mag, there are a few objects with only 2MASS or MKO photometry that changes the input sample slightly between 2MASS and MKO.) A subset of 76 of these objects had magnitudes of MH < 13.3 mag, which we used in our H-band relations.

      We fit luminosity as a function of K-band absolute magnitude with a fifth-order polynomial, since there was a clear structure in the residuals using only a third-order polynomial. For the more limited H-band magnitude range, a third-order polynomial was sufficient. In the K-band, it was visually apparent that the scatter at brighter magnitudes was less than at fainter magnitudes. The fact that the nominal errors on the magnitudes and luminosities are not significantly larger for the fainter objects implies that this increased scatter is due to either unquantified systematic errors in the luminosities of the lower luminosity objects or real astrophysical scatter in the relationship between the K-band absolute magnitude and luminosity. Regardless, we quantify this effect by separately computing the rms scatter at the bright end (MK ≤ 13.0 mag; rms of 0.04 dex for 90 objects) and faint end (rms of 0.07 dex for 36 objects). The coefficients for all relations are given in Table 14, and the fits and residuals are displayed in Figure 21.

      Figure 21.

      Figure 21. Polynomial relations between bolometric luminosity and absolute magnitude (red) derived from input data from Filippazzo et al. (2015; black points) as described in Appendix A.3. The bottom panels show the residuals in the input data after subtracting the polynomials. We use these relations to derive the component luminosities for our sample binaries. When both components have MH < 13.3 mag, we use the H-band relations, otherwise we use K-band relations.

      Standard image High-resolution image

      Appendix B: Discussion of Individual Objects

      B.1. LP 349-25AB (M7+M8)

      In our previous work, we used a parallax of 75.8 ± 1.6 mas from Gatewood & Coban (2009) to compute luminosities and a total mass that gave a Lyon Dusty model-derived age of ${127}_{-17}^{+21}$ Myr (Dupuy et al. 2010). Our new more accurate parallax of 69.2 ± 0.9 mas from CFHT results in somewhat higher masses and luminosities but still the youngest model-derived age of any object in our sample, ${271}_{-29}^{+22}$ Myr (BHAC). We now have directly measured individual masses, and the ages derived from the component masses and luminosities are coeval within the uncertainties (Δt = 50 ± 50 Myr). Our measured mass ratio ($q={0.941}_{-0.030}^{+0.029}$) is consistent within the uncertainties with the model-derived mass ratio from our total-mass analysis (${0.88}_{-0.04}^{+0.03}$). Such a young age implies that LP 349-25AB is a pair of pre-main-sequence stars with masses of 85 ± 4 ${M}_{\mathrm{Jup}}$ and 80 ± 3 ${M}_{\mathrm{Jup}}$. At a distance of only ${14.45}_{-0.19}^{+0.18}$ pc, this is the nearest pre-main-sequence system containing very low-mass stars (<0.1 M), with the next closest being LSPM J1314+1320AB (M7+M7; Dupuy et al. 2016) at 17.25 pc.

      According to the BHAC tracks, a 0.075 M object comparable to LP 349-25B requires ≈130 Myr to deplete 99.9% of its initial supply lithium, while a 0.075 M object comparable to LP 349-25B would need ≈110 Myr. Therefore, BHAC models predict both components are fully depleted in lithium, in agreement with the nondetection of Li i absorption (<0.5 Å) reported by Reiners & Basri (2009). LP 349-25AB therefore provides a critical test of the lithium depletion boundary in low-mass stars and brown dwarfs at the young ages used to age-date young clusters at ∼20–200 Myr (e.g., Bildsten et al. 1997; Binks & Jeffries 2014; Kraus et al. 2014). It joins LSPM J1314+1320AB as the only other pre-main-sequence system where objects of known mass empirically constrain the model-predicted lithium depletion boundary (Dupuy et al. 2016). Like LSPM J1314+1320AB, LP 349-25AB is a luminous radio source (Phan-Bao et al. 2007; Osten et al. 2009), which makes it a rare benchmark for characterizing radio emission in ultracool dwarfs. High-precision VLBI astrometry may be possible for LP 349-25AB, like Forbrich et al. (2016) recently showed for LSPM J1314+1320AB, which would refine the individual mass precision and enable even stronger constraints on models.

      LP 349-25AB has not previously been reported as having spectral signatures of low surface gravity. However, examination of our own integrated light SpeX SXD spectrum (R ≈ 1200) indicates that it has a classification of M8 int-g on the Allers & Liu (2013) system. Given the gap in known young associations between the ages of the AB Doradus (∼150 Myr) and Ursa Majoris (∼400 Myr) groups, LP 349-25AB (250–300 Myr) provides unique evidence that spectral signatures of low gravity in the infrared can persist to at least 250 Myr for late-M dwarfs. There is only one other system of ultracool dwarfs with spectral signatures of youth that have directly measured masses, LSPM J1314+1320AB (92.8 ± 0.6 ${M}_{\mathrm{Jup}}$ and 91.7 ± 1.0 ${M}_{\mathrm{Jup}}$; Dupuy et al. 2016). Despite LSPM J1314+1320AB having more marginal signatures of low gravity, its components actually have lower model-derived surface gravities ($\mathrm{log}(g)\simeq 4.8$ dex) than the older and somewhat less massive LP 349-25AB (5.1 dex). This implies that while the current gravity classification system does indeed pick out young objects (≲300 Myr), there is a limit to the granularity by which relative strengths of spectral features can be used to distinguish surface gravity and thus age.

      B.2. LP 415-20AB (M6+M8)

      LP 415-20AB has the largest total mass in our sample (${248}_{-29}^{+26}$ ${M}_{\mathrm{Jup}}$) and also the largest fractional uncertainty in its mass (${}_{-12}^{+10} \% ).$ Given the component spectral types of M6.0 ± 1.0 and M8.0 ± 0.5, and the integrated-light optical type of M7.5 found by Gizis et al. (2000a), a mass as large as 250 ${M}_{\mathrm{Jup}}$ would be quite puzzling. The implied component masses would be ≈120–130 ${M}_{\mathrm{Jup}}$, and objects that massive are not known at such late spectral types (e.g., L726-8AB and L789-6ABC are both M5.5 in integrated light; Delfosse et al. 2000). The size of our mass uncertainty suggests that LP 415-20AB could simply be a ≳2σ outlier, but we also consider the possibility that it is an unresolved triple system. Our directly measured mass ratio of ${0.59}_{-0.12}^{+0.10}$ is rather far from unity despite a modest luminosity ratio of ${\rm{\Delta }}\mathrm{log}({L}_{\mathrm{bol}}/{L}_{\odot })\,={0.25}_{-0.03}^{+0.04}$ dex. This is also consistent with higher-order multiplicity, but the mass ratio uncertainty is too large to make a definitive statement.

      In our total-mass analysis, rejection sampling eschewed masses at the high end of the input distribution, resulting in ${M}_{\mathrm{tot}}={201}_{-3}^{+4}$ ${M}_{\mathrm{Jup}}$, 1.7σ lower than our measurement (Table 13). This is because the models do not allow the higher masses to correspond to the luminosity we measure for LP 415-20A (Figure 11). The individual-mass analysis confirms that this mass discrepancy is found to be due entirely to the primary. The mass of LP 415-20A after rejection sampling is ${109.8}_{-2.9}^{+3.1}$ ${M}_{\mathrm{Jup}}$, 2.6σ lower than the input ${156}_{-18}^{+17}$ ${M}_{\mathrm{Jup}}$, whereas the input and output secondary masses are essentially unchanged. Therefore, both sets of analyses suggest that if our unusually high mass is due to an unresolved component of LP 415-20A, its mass is expected to be ≈50 ${M}_{\mathrm{Jup}}$ (assuming that the unresolved component contributes negligibly to the luminosity).

      As a test, we also performed our analysis for a hypothetical scenario where LP 415-20A is an equal-mass, equal-flux binary with each component having $M={78}_{-9}^{+8}$ ${M}_{\mathrm{Jup}}$ and $\mathrm{log}({L}_{\mathrm{bol}}/{L}_{\odot })=-3.32\pm 0.03$ dex. This scenario is the simplest to implement because we do not need to invoke a mass–luminosity relation to apportion flux between the hypothetical components. We found reasonable agreement with models, though the resulting LP 415-20A component masses after rejection sampling preferred a somewhat higher mass of ${92.3}_{-1.6}^{+2.2}$ ${M}_{\mathrm{Jup}}$ (Table 13). This is consistent with the idea that if LP 415-20A is an unresolved binary, its components are somewhat unequal in mass. However, we reiterate that this discrepancy could also simply be a statistical outlier, e.g., due to the somewhat uncertain parallax (σπ/π = 3%) that dominates the mass uncertainty. As one of the brighter targets in our sample, LP 415-20AB should be well detected by Gaia, which would resolve the current mass discrepancy.

      In all cases above, the masses and luminosities of the components of LP 415-20 are consistent with being coeval stars, most likely on the main sequence but also consistent with ages of a few hundred Myr (the BHAC model-derived age for LP 415-20B is ${5.0}_{-4.7}^{+1.9}$ Gyr). Thus, from our model-derived age distribution alone, we do not rule out the possibility that LP 415-20 could be a member of the Hyades (750 ± 150 Myr; Brandt & Huang 2015), as was originally noted by Bryja et al. (1992). We can however use the kinematic information, both from our own astrometry and the literature, to reassess its potential membership. We have projected the space velocity of the Hyades, $(U,V,W)=(42.3,-19.1,-1.5)$ km s−1 (de Bruijne et al. 2001), into proper motion and radial velocity using our R.A., decl., and parallax of LP 415-20. We simulate the 0.3 km s−1 velocity dispersion of the Hyades and our parallax uncertainty in a Monte Carlo fashion, finding ${\mu }_{{\alpha }^{* }}=132.2\pm 4.4$ mas yr−1, ${\mu }_{\delta }=-42.8\pm 2.1$ mas yr−1, and ${v}_{\mathrm{rad}}=38.8\pm 0.3$ km s−1. In comparison, our measured absolute proper motion is 126.1 ± 0.7 mas yr−1 and −38.2 ± 0.8 mas yr−1, which is only 8.5 ± 3.4 mas yr−1 (1.6 ± 0.5 km s−1) away from the predicted proper motion assuming Hyades kinematics. Thus, under these assumptions, the motion of LP 415-20AB appears fairly inconsistent (2.5σ) with membership. The mean radial velocity of the two components from Konopacky et al. (2010) is 40.8 ± 1.4 km s−1, only 2.0 ± 1.4 km s−1 different from the predicted Hyades value. The full 3D distance in UVW between the proper motion and RV measurements and the predicted Hyades values is 2.7 ± 1.1 km s−1, indicating that it is only a marginal candidate for membership in the Hyades. However, if the velocity dispersion of the Hyades were larger than currently thought, e.g., 1 km s−1 as is seen for other young clusters, then LP 415-20AB would be a stronger candidate member.

      B.3. SDSS J0423−0414AB (L6.5+T2)

      SDSS J0423−0414AB has the lowest total mass in our sample (83 ± 3 ${M}_{\mathrm{Jup}}$), and it also has a well-determined mass ratio from our CFHT data (0.62 ± 0.04), yielding individual masses of ${51.6}_{-2.5}^{+2.3}$ ${M}_{\mathrm{Jup}}$ and ${31.8}_{-1.6}^{+1.5}$ ${M}_{\mathrm{Jup}}$. Our total-mass analysis gives an SM08 model-derived mass ratio of ${0.65}_{-0.06}^{+0.05}$, in good agreement with our measurement, and a system age of ${0.81}_{-0.09}^{+0.07}$ Gyr. In contrast, the Lyon Cond models give a mass ratio of 0.80 ± 0.05 (2.8σ discrepant with our measurement) but a very similar age (${0.80}_{-0.08}^{+0.07}$ Gyr). Our measured mass ratio of 0.62 ± 0.04 is much lower than previous estimates, e.g., 0.8–1.0 from Burgasser et al. (2005) and 0.80 ± 0.03 from Liu et al. (2010).

      The case of SDSS J0423−0414AB is similar to SDSS J1052+4422AB (L6.5 ± 1.5+T1.5 ± 1.0, SM08 age of ${1.04}_{-0.15}^{+0.14}$ Gyr), with a nearly identical primary mass but with mass and luminosity ratios further from unity. Like what we find for SDSS J0423−0414AB, Dupuy et al. (2015b) found a similar discrepancy between the measured and the Lyon model-derived mass ratios for SDSS J1052+4422AB, but in agreement with the SM08 models. Using the individual mass and luminosity of each component independently, in both cases the Lyon models predict a younger age for the secondary as compared to the primary, Δt = −0.44 ± 0.15 Gyr for SDSS J0423−0414AB and $-{0.35}_{-0.22}^{+0.25}$ Gyr for SDSS J1052+4422AB. This is caused by the secondary being more luminous than expected according to Lyon models, resulting in a younger model-derived age. In contrast, the SM08 models give a higher luminosity for the secondary because they predict that luminosity does not fade as quickly during and immediately following the transition from cloudy to cloud-free atmospheres, which they prescribe ad hoc to occur as the temperature drops from 1400 K to 1200 K. The components of SDSS J0423−0414AB have SM08 model-derived temperatures straddling this transition, ${1430}_{-40}^{+30}$ K and 1200 ± 40 K.

      Both components of SDSS J0423−0414AB have masses so low that they should never significantly deplete their initial supply of lithium (≲55 ${M}_{\mathrm{Jup}}$; Chabrier & Baraffe 2000). However, the monatomic lithium that is readily probed by the Li i doublet at 6708 Å can be removed from the atmosphere chemically. Models from Lodders (1999) indicate that monatomic lithium ceases to be the dominant lithium-bearing species below ${T}_{\mathrm{eff}}\approx 1500$ K at pressures near the photosphere (∼1 bar). Detailed spectral synthesis modeling shows that Li i absorption can in fact persist to much lower temperatures (e.g., Pavlenko et al. 2000; Allard et al. 2001), and indeed it has been detected in objects as cool as WISE J1049−5319B (T0; Faherty et al. 2014; Lodieu et al. 2015). Kirkpatrick et al. (2008) reported strong Li i absorption (EW = 11 Å) in the integrated-light spectrum of SDSS J0423−0414AB. This is consistent with our Lyon Cond model analysis that predicts SDSS J0423−0414A should still possess all or most of its initial lithium, $\mathrm{log}(\mathrm{Li}/{\mathrm{Li}}_{\mathrm{init}})=-{0.04}_{-0.20}^{+0.04}$ dex, along with the previous detection of lithium in even cooler objects than SDSS J0423−0414A.

      B.4. 2MASS J0700+3157AB (L3+L6.5)

      Our measured mass ratio of q = 1.08 ± 0.05 indicates that 2MASS J0700+3157B is in fact slightly more massive than 2MASS J0700+3157A, despite the fact that it is 0.50 ± 0.06 dex less luminous. Combined with our total mass of ${141}_{-5}^{+4}$ ${M}_{\mathrm{Jup}}$, we find individual masses for 2MASS J0700+3157A and 2MASS J0700+3157B of 68.0 ± 2.6 ${M}_{\mathrm{Jup}}$ and ${73.0}_{-3.0}^{+2.9}$ ${M}_{\mathrm{Jup}}$, respectively. These masses and luminosities result in very different SM08 model-derived ages, ${0.76}_{-0.14}^{+0.09}$ Gyr for 2MASS J0700+3157A and ${7.8}_{-5.2}^{+2.8}$ Gyr for 2MASS J0700+3157B, or in other words, a difference of ${\rm{\Delta }}\mathrm{log}t={1.00}_{-0.20}^{+0.29}$ dex.

      The most natural explanation is that 2MASS J0700+3157B is an unresolved binary itself composed of two brown dwarfs. We performed a simple test, dividing the mass and luminosity of 2MASS J0700+3157B in two (36.7 ± 1.5 ${M}_{\mathrm{Jup}}$, −4.75 ± 0.04 dex) and repeating our model analysis. In this hypothetical scenario, we found model-derived ages that agreed better between 2MASS J0700+3157A and 2MASS J0700+3157B (${\rm{\Delta }}\mathrm{log}t\,={0.18}_{-0.07}^{+0.09}$ dex for SM08, $-{0.09}_{-0.08}^{+0.09}$ dex for Cond).

      Interestingly, the total mass of 2MASS J0700+3157AB is not high enough to betray its unresolved higher-order multiplicity without additional mass-ratio information. When we performed our total-mass analysis, evolutionary models are able to self-consistently apportion the total mass between the two components at their appropriate luminosities at a single coeval age (${1.9}_{-0.7}^{+0.6}$ Gyr for SM08, ${2.1}_{-0.6}^{+0.4}$ Gyr for Cond). But the model-derived mass ratio, e.g., ${0.86}_{-0.03}^{0.04}$ from Cond, is 3.4σ lower than what we measure. This case highlights the important role directly measured mass ratios play in the substellar regime, where a given mass does not correspond to a particular luminosity or spectral type unlike for main-sequence stars.

      Thorstensen & Kirkpatrick (2003) reported a lack of Li i absorption in the integrated-light spectrum of 2MASS J0700+3157AB (EW < 0.3 Å). This is consistent with the prediction from the Lyon models that the 68.0 ± 2.6 ${M}_{\mathrm{Jup}}$ primary should deplete 99.9% of its lithium within ≈200 Myr. Given the unknown nature of the unresolved secondary's individual components, it is not clear whether or not they should cause lithium absorption to the integrated-light spectrum. When present, lithium is routinely detected at EW ∼ 10 Å in late-L dwarfs, so it seems plausible that if one of the unresolved components of 2MASS J0700+3157B were lithium bearing, it would have been detectable even when diluted by up to a factor of ∼20 in flux given typical detection limits of EW ∼ 0.5 Å. The lack of a Li i detection therefore hints at two possible scenarios for the unresolved components. (1) A lithium-depleted higher mass object is paired with a lower mass object that either has lithium but is too diluted in flux or has lithium in molecular form only (LiCl, LiOH) because it is very cool. (2) 2MASS J0700+3157B is composed of two cool, chemically depleted objects. In the equal-flux, equal-mass scenario, the absolute magnitudes of the components would be MJ = 15.06 ± 0.04 mag and MK = 13.44 ± 0.04 mag (MKO), roughly consistent with the faint end of field objects with spectral type L7–T0 (see Table 15 of Dupuy & Liu 2012). This is the same range of spectral type over which lithium becomes undetectable in field objects, with later-type objects being more likely to be chemically depleted in lithium. Later-type objects would also display some methane absorption, and this could be detectable in the spectrum of 2MASS J0700+3157B even if it were somewhat diluted by an earlier-type component. No evidence of methane is seen in the integrated-light spectrum of 2MASS J0700+3157 (see Figure 13 of Dupuy & Liu 2012), so resolved spectroscopy with AO or spectro-astrometry will be needed to further examine the possible presence of an unresolved methane-bearing component.

      2MASS J0700+3157 was originally discovered by Thorstensen & Kirkpatrick (2003) as an unresolved single L3.5 dwarf displaying parallax and proper motion during the course of an astrometry program targeting cataclysmic variables. Therefore, unlike most other known ultracool dwarfs, the discovery of 2MASS J0700+3157 did not involve a traditional magnitude-selected sample, which would have been subject to a Malmquist bias favoring the discovery of such high-order multiples. However, our own dynamical-mass sample is more complete for the most massive systems as they orbit faster at a given semimajor axis ($P\propto {{M}_{\mathrm{tot}}}^{-1/2}$), and our orbit determinations are mostly limited by having sufficient time baseline. Therefore, our dynamical-mass sample is slightly biased toward more massive systems. As an example of the potential amplitude of such bias, the period we measure for 2MASS J0700+3157AB is 23.9 ± 0.5 years, but if this system contained only two-thirds of the mass, the period would have been 29.3 years.

      B.5. LHS 1901AB (M7+M7)

      In our previous work, we used a parallax of 77.8 ± 3.0 mas from Lépine et al. (2009) to compute a total mass of ${203}_{-22}^{+26}$ ${M}_{\mathrm{Jup}}$ (Dupuy et al. 2010). Our new more accurate parallax from CFHT (76.4 ± 1.1 mas) and updated orbit determination results in a consistent but more precise mass of ${213}_{-10}^{+9}$ ${M}_{\mathrm{Jup}}$. We also now have directly measured individual masses of 113 ± 8 ${M}_{\mathrm{Jup}}$ and 99 ± 7 ${M}_{\mathrm{Jup}}$, based on our mass ratio of ${0.87}_{-0.11}^{+0.09}$. In our total-mass analysis, the BHAC model-derived mass ratio is somewhat higher but consistent within the errors (${0.97}_{-0.04}^{+0.03}$), as expected given that the luminosity ratio is near unity (${\rm{\Delta }}\mathrm{log}({L}_{\mathrm{bol}})=0.04\pm 0.03$ dex). Given our measured masses and luminosities, LHS 1901AB is consistent with being a main-sequence stellar binary, with a correspondingly unconstrained age (BHAC models only give a 3σ lower limit of >0.3 Gyr from our total mass analysis). As noted by Dupuy et al. (2010), this system is likely to be old given that it is lacking Hα emission, which is unusual for late-M dwarfs (e.g., West et al. 2008). Finally, we note that in our latest CFHT imaging from 2016, which we do not use here, LHS 1901AB has become blended with a background source that had previously been well-separated, making these images unusable for absolute astrometry.

      B.6. 2MASS J0746+2000AB (L0+L1.5)

      The masses of 2MASS J0746+2000A (${82.4}_{-1.5}^{+1.4}$ ${M}_{\mathrm{Jup}}$) and 2MASS J0746+2000B (78.4 ± 1.4 ${M}_{\mathrm{Jup}}$) have the highest fractional precision (2%) of any individual mass in our sample. Given our determination of the substellar boundary of ≈70 ${M}_{\mathrm{Jup}}$, the masses indicate that this binary is a pair of stars and that neither is a brown dwarf, in contrast to the analysis of Bouy et al. (2004) but in agreement with Gizis & Reid (2006). Our measured mass ratio of ${0.952}_{-0.027}^{+0.026}$ agrees well with the BHAC model-derived mass ratio from our total mass analysis (${0.957}_{-0.010}^{+0.012}$), indicating that our masses and luminosities follow BHAC isochrones well. However, because the components of 2MASS J0746+2000AB are both main-sequence stars, the age is essentially unconstrained by models, which only give 3σ lower limits of >0.9 Gyr for 2MASS J0746+2000A and >1.0 Gyr for 2MASS J0746+2000B.

      One of the components of 2MASS J0746+2000AB is a radio emitter with a rotation period of 2.0720 ± 0.0018 hr (Berger et al. 2009). In addition, Harding et al. (2013) report optical variability in integrated light with a period of 3.32 ± 0.15 hr. Unfortunately, it is not known which component corresponds to which period. Konopacky et al. (2012) report $v\sin {i}_{\star }=19\pm 2$ km s−1 and 33 ± 3 km s−1 for the primary and secondary components, respectively, suggesting that the secondary may be a radio emitter. We consider this possibility (Case I) and the alternative (Case II) and compare the implied stellar radii from combining the rotation periods with projected rotational velocities to the radii predicted from evolutionary models. In Case I, we compute minimum radii of the primary and secondary of $R\sin {i}_{\star }=0.50\pm 0.06$ ${R}_{\mathrm{Jup}}$ and 0.55 ± 0.05 ${R}_{\mathrm{Jup}}$, respectively. In Case II we compute minimum radii of the primary and secondary of $R\sin {i}_{\star }=0.32\pm 0.03$ ${R}_{\mathrm{Jup}}$ and 0.89 ± 0.09 ${R}_{\mathrm{Jup}}$, respectively.

      BHAC models predict radii of ${0.959}_{-0.008}^{+0.007}$ ${R}_{\mathrm{Jup}}$ and ${0.914}_{-0.008}^{+0.007}$ ${R}_{\mathrm{Jup}}$ for the primary and secondary, respectively. Both possibilities that we consider are consistent with these model radii, i.e., no minimum radii are larger than these values, but they would correspond to very different projected alignments of the stellar spin axes. In Case I, inclinations of i = 32 ± 4° and 37 ± 4° for the primary and secondary, respectively, would be needed to match the model radii. This is consistent with the co-alignment of stellar spin axes, as well as rough alignment with the orbital plane ($i=138\buildrel{\circ}\over{.} {56}_{-0.21}^{+0.20}$ means that the spin axes would need to be i = 138fdg56 or 41fdg44 to be aligned). In contrast, for Case II, projected stellar inclinations of i = 19° ± 2° and 67° ± 4° for the primary and secondary, respectively, would be needed to match the model radii. This would rule out either of the stellar spin axes from being aligned with each other or with the orbital plane. We conclude that Case I is more likely because of the remarkable coincidence of the stellar spin axes and orbital inclination all being within ±5° of each other. However, we note that this is still only a probabilistic argument and that the stellar spin axis angles are only seen in projection, and we cannot rule out or determine if they are aligned or misaligned out of the plane of the sky.

      B.7. 2MASS J0920+3517AB (L5.5+L9)

      The integrated-light spectral type of 2MASS J0920+3517AB is L6.5 in the optical (Kirkpatrick et al. 2000) and T0p in the infrared (Burgasser et al. 2006a). We determined component types of L5.5 ± 1.0 and L9.0 ± 1.5 from spectral deconvolution (Dupuy & Liu 2012). It is therefore surprising that we find a mass of ${116}_{-8}^{+7}$ ${M}_{\mathrm{Jup}}$ for the fainter 2MASS J0920+3517B, much higher than the substellar boundary (≈70 ${M}_{\mathrm{Jup}}$). Moreover, the total system mass 187 ± 11 ${M}_{\mathrm{Jup}}$ suggests that there is more mass in this system than two brown dwarfs. The most plausible explanation is that 2MASS J0920+3517B is itself an unresolved binary.

      Using only the primary mass (71 ± 5 ${M}_{\mathrm{Jup}}$) and luminosity, SM08 models give an essentially unconstrained age of ${7}_{-5}^{+3}$ Gyr and Cond models give a somewhat more tightly constrained age of ${3.1}_{-1.7}^{+1.5}$ Gyr. If we arbitrarily divide 2MASS J0920+3517B into two equal-mass (${58}_{-4}^{+3}$ ${M}_{\mathrm{Jup}}$), equal-luminosity components, then we find ages from SM08 (${2.3}_{-0.4}^{+0.3}$ Gyr) and Cond (${1.99}_{-0.37}^{+0.25}$ Gyr) that are in reasonable agreement with the primary. Therefore, the model-derived ages are consistent with the hypothesis that 2MASS J0920+3517B is an unresolved, nearly equal-mass binary.

      The nondetection of Li i absorption in the integrated light spectrum (<0.5 Å; Kirkpatrick et al. 2000) is consistent with the fact that the primary mass (71 ± 5 ${M}_{\mathrm{Jup}}$) is well above the 55–60 ${M}_{\mathrm{Jup}}$ lithium depletion boundary at field ages (Section 7.2). In principle, a lithium-bearing component in 2MASS J0920+3517B could still be detectable even when diluted by the somewhat brighter 2MASS J0920+3517A (${\rm{\Delta }}F814W=0.30\pm 0.10$ mag). In our hypothetical equal-mass binary scenario for 2MASS J0920+3517B, our interpolation of Cond models predicts only modest lithium depletion of $\mathrm{log}(\mathrm{Li}/{\mathrm{Li}}_{\mathrm{init}})=-{0.0215}_{-0.0018}^{+0.0019}$ dex for the components. The case of 2MASS J0920+3517B is therefore quite similar to that of 2MASS J0700+3157B, which we discussed in detail above. We note that although the unresolved components of 2MASS J0920+3517B are likely both substellar, the primary 2MASS J0920+3517A could be a star or brown dwarf within the measurement uncertainties.

      B.8. 2MASS J1017+1308AB (L1.5+L3)

      2MASS J1017+1308AB (L1.5+L3) has one of the larger fractional uncertainties in total mass (${}_{-12}^{+9} \% )$ in our sample, owing to a somewhat large parallax uncertainty (3.4%). Our directly measured mass ratio (0.92 ± 0.08) results in component masses of ${81}_{-11}^{+10}$ ${M}_{\mathrm{Jup}}$ and 75 ± 7 ${M}_{\mathrm{Jup}}$. Our measured mass ratio agrees well with the model-derived mass ratios from our total mass analysis (${0.987}_{-0.015}^{+0.017}$ from BHAC, and SM08 gives essentially the same result). This means that our individual masses and luminosities agree with both model mass–luminosity relations, although this system does not offer a very strong test of models given the near-unity mass and luminosity ratios. Because the components of 2MASS J1017+1308AB are both likely main-sequence stars, the system age is essentially unconstrained by models, which only give 3σ lower limits of >0.6 Gyr (BHAC) and >0.5 Gyr (SM08) from our total-mass analysis.

      B.9. 2MASS J1047+4026AB a.k.a. LP 213-68AB (M8+L0)

      We measure a total mass of ${178}_{-12}^{+11}$ ${M}_{\mathrm{Jup}}$ for 2MASS J1047+4026AB and individual masses of ${97}_{-7}^{+6}$ ${M}_{\mathrm{Jup}}$ and 80 ± 6 ${M}_{\mathrm{Jup}}$. Our measured mass ratio of 0.82 ± 0.06 is somewhat lower than but consistent within the uncertainties of the BHAC model-derived mass ratio from our total-mass analysis (0.947 ± 0.029), indicating that our masses and luminosities are consistent with BHAC isochrones. However, because the components of 2MASS J1047+4026AB are both main-sequence stars, the age is essentially unconstrained by models, which only give 3σ lower limits of >0.5 Gyr from our total-mass analysis.

      The M8 dwarf 2MASSW J1047138+402649 (a.k.a. LP 213-68) is a proper motion companion to the M6.5 dwarf LP 213-67 (2MASSW J1047126+402643, 14farcs4 away) as described by Gizis et al. (2000b) in their paper presenting the discovery of both objects as ultracool dwarfs. Close et al. (2003) reported the discovery of several binaries from their Gemini AO survey, but they report conflicting information regarding this source. Close et al. (2003) use the 2MASS name of the M6.5 dwarf LP 213-67 but refer to the object they observed as an M8 dwarf. We believe that they indeed observed the M8 dwarf LP 213-68, given that we have observed more than a full orbital period of this binary with Keck and the separation and P.A. reported by Close et al. (2003) agree very well with our orbit. If Close et al. (2003) observed LP 213-67, then it would have to be a binary that happened to have a separation and P.A. matching our ephemeris for LP 213-68 at the epoch of their observations, which would be a remarkable coincidence. Bouy et al. (2008) present Lick AO observations with very clear details indicating that they observed LP 213-68 (despite referring to it by the 2MASS name of LP 213-67 in their tables), and their reported separation and P.A. (106 ± 14 mas and 319.3 ± 1fdg0) agree reasonably well with our ephemeris for LP 213-68 at that epoch (109 ± 2 mas and 307.3 ± 1fdg9). In contrast, Konopacky et al. (2010) report observations that do not appear to be consistent with our ephemeris of LP 213-68. They detected a binary at 32 ± 2 mas on 2006 June 21 UT but nothing outside of 47 mas on 2006 November 27 UT and 2007 December 2 UT. Our ephemeris predicts separations and P.A.s on those respective dates of 87 ± 5 mas and 292° ± 3°; 101 ± 3 mas and 302° ± 2°; and 117.5 ± 1.0 mas and 316fdg6 ± 1fdg6. It is possible that Konopacky et al. (2010) observed the M6.5 dwarf LP 213-67 instead of the M8 dwarf LP 213-68 because their tables use the 2MASS name and coordinates of LP 213-67. We conclude that this is the most likely explanation for their measurements. Interestingly, this implies that the M6.5 companion is also a binary, even tighter than LP 213-68, making this a hierarchical quadruple system. Unlike other known quadruples containing ultracool dwarfs (e.g., Gl 337, 2MASS J04414565+2301580) the LP 213-67/LP 213-68 system seems to be entirely composed of >M6 dwarfs.

      B.10. SDSS J1052+4422AB (L6.5+T1.5)

      The properties of SDSS J1052+4422AB are discussed in detail in Dupuy et al. (2015b). Our new homogeneous analysis here results in slightly revised properties for both mass and luminosity, but the key results are unchanged. Our measured mass ratio of 0.78 ± 0.07 is further from unity than expected given the near-unity luminosity ratio ${\rm{\Delta }}\mathrm{log}({L}_{\mathrm{bol}})=0.13\pm 0.08$ dex. The mass ratio derived from the SM08 models (${0.82}_{-0.11}^{+0.09}$, 0.3σ different) provides a better match to our measurement than that from Cond models (0.91 ± 0.05, 1.5σ different). Or put in terms of age, our directly measured individual masses and luminosities yield ages that agree within the errors with the SM08 models (${\rm{\Delta }}\mathrm{log}t=-0.02\pm 0.12$ dex) but somewhat discrepant ages with the Cond models (${\rm{\Delta }}\mathrm{log}t=-0.15\pm 0.10$ dex).

      B.11. Gl 417BC (L4.5+L6)

      The properties of Gl 417BC and the corresponding astrophysical interpretation are discussed in detail in Dupuy et al. (2014). Our new homogeneous analysis here results in slightly revised properties for both mass and luminosity, but the key results are unchanged.

      Gl 417 is the only system in our sample for which a Gaia DR1 parallax is available from the Tycho-Gaia Astrometric Solution (Lindegren et al. 2016). Unfortunately, this parallax (43.86 ± 0.34 mas) is 3.1σ discrepant with the Hipparcos value (45.61 ± 0.44 mas) used in our previous work and the analysis presented here. If we adopt the Gaia-DR1 parallax, the total mass would be higher (${111.6}_{-3.1}^{+2.9}$ ${M}_{\mathrm{Jup}}$, compared to ${99.2}_{-3.3}^{+3.0}$ ${M}_{\mathrm{Jup}}$), the component luminosities would be 0.03 dex higher, and the model-derived ages would be correspondingly older. Using the Hipparcos parallax, both the BHAC and SM08 models give an age of ${490}_{-40}^{+30}$ Myr, but using the Gaia-DR1 parallax, the BHAC and SM08 models would give ages of 620 ± 40 Myr and 600 ± 40 Myr, respectively. Such older ages are more consistent with the gyrochronology-derived age of ${740}_{-120}^{+150}$ Myr for the primary star Gl 417A and would consequently reduce the significance of the substellar "luminosity problem" for this system. The quality of the Gaia-DR1 parallaxes have only just begun to be assessed by the community, and we defer judgment on the discrepancy for this system until the Tycho-Gaia Astrometric Solution errors are better understood.

      B.12. LHS 2397aAB (M8+L dwarf)

      We originally presented an analysis of this system based on the total mass alone in Dupuy et al. (2009c), but we noted that the photocenter orbit should be readily measurable given a long-enough time baseline. Indeed, we detect the photocenter motion of LHS 2397aAB at the highest significance of any binary in our sample (37σ), allowing us to measure a mass ratio of ${0.706}_{-0.028}^{+0.027}$. Only the Cond models cover the mass and luminosity of both the primary and secondary, and they predict a mass ratio (${0.75}_{-0.04}^{+0.07}$) that is in excellent agreement with our much more precise measurement. As expected, the primary is a star (93 ± 4 ${M}_{\mathrm{Jup}}$), and the secondary is consistent with being substellar (66 ± 4 ${M}_{\mathrm{Jup}}$). The Cond model-derived age of the secondary is ${2.6}_{-1.0}^{+0.6}$ Gyr, while the age of the primary is essentially unconstrained but consistent with the secondary. The SM08 models give a similar age for the secondary of ${2.8}_{-1.5}^{+2.1}$ Gyr.

      Despite the fact that the components of LHS 2397aAB are different in luminosity by 1.14 ± 0.06 dex, further from unity than any other binary in our sample, the fact that the components are a star and a brown dwarf means that this system on its own does not actually provide a strong test of the mass–luminosity relation predicted by models. Such a wide range of ages are allowed for the primary given that it is on the main sequence that models are free to match the mass and luminosity of the secondary exactly by fine-tuning the age. Models predict that LHS 2397aB is strongly depleted in lithium, $\mathrm{log}(\mathrm{Li}/{\mathrm{Li}}_{\mathrm{init}})=-{1.8}_{-0.6}^{+0.7}$ dex, which is consistent with its nondetection (EW < 0.5 Å) in the integrated-light spectrum from Reiners & Basri (2009).

      B.13. Kelu-1AB (L2+L4)

      Kelu-1AB is the only one of our sample binaries with a CFHT parallax but without a measured mass ratio, owing to the relatively short time baseline (2.26 years) of the available CFHT observations. Moreover, our parallax (48.0 ± 2.2 mas) disagrees somewhat with other comparably precise published values of 53.6 ± 2.0 mas (Dahn et al. 2002) and 51.75 ± 1.16 mas (Weinberger et al. 2016). Using our parallax, the total mass is ${180}_{-26}^{+22}$ ${M}_{\mathrm{Jup}}$. In comparison, the Dahn et al. (2002) parallax gives ${129}_{-16}^{+13}$ ${M}_{\mathrm{Jup}}$ and the Weinberger et al. (2016) parallax gives ${144}_{-10}^{+9}$ ${M}_{\mathrm{Jup}}$. Given the lack of mass ratio information and that the published parallaxes, including our own, span such a wide range of possible masses, we exclude Kelu-1AB from our analysis. The fact that Li i absorption is detected in integrated light (EW = 1.7 Å; Kirkpatrick et al. 1999) implies that at least one component in the system should be ≲60 ${M}_{\mathrm{Jup}}$ according to BHAC models, which would potentially favor the lower system mass (larger parallax). Liu & Leggett (2005) noted that the components of Kelu-1AB might straddle the lithium fusion boundary, providing a uniquely constraining test of models. However, the current uncertainty in the dynamical mass is too large to distinguish whether the components of Kelu-1AB are both stars or brown dwarfs, let alone assess whether they are lithium bearing.

      In an unrefereed work, Stumpf et al. (2008) proposed that Kelu-1AB is a triple system. They computed a dynamical mass with large errors (${185}_{-58}^{+118}$ ${M}_{\mathrm{Jup}}$)16 and found an H-band feature in their resolved spectrum of Kelu-1A that they claimed as evidence of an unresolved T-dwarf companion. This feature is strong enough that it should appear in the integrated-light spectrum of the whole system even after being diluted by Kelu-1B, but an inspection of the published unresolved spectrum of Kelu-1AB shows no such feature. We therefore conclude that there is no strong evidence that Kelu-1AB is actually a triple system.

      B.14. 2MASS J1404−3159AB (L9+T5)

      2MASS J1404−3159AB is a J-band flux-reversal binary (Looper et al. 2008), and using our photometry here we find component spectral types of L9.0 ± 1.0 and T5.0 ± 0.5 by spectral decomposition, making 2MASS J1404−3159B one of the latest-type objects in our sample. We have a well-determined total mass (${120}_{-13}^{+11}$ ${M}_{\mathrm{Jup}}$) and mass ratio (0.84 ± 0.06), resulting in individual masses of 65 ± 6 ${M}_{\mathrm{Jup}}$ and ${55}_{-7}^{+6}$ ${M}_{\mathrm{Jup}}$. In our total-mass analysis, both sets of models give mass ratios in good agreement with our measurement (${0.79}_{-0.11}^{+0.08}$ from SM08 and ${0.85}_{-0.06}^{+0.07}$ from Cond). The model-derived ages are also consistent, with SM08 giving ${3.0}_{-1.3}^{+0.8}$ Gyr and Cond giving ${2.5}_{-0.9}^{+0.6}$ Gyr.

      The only other J-band flux-reversal binary in our final mass sample is SDSS J1052+4422AB, and its mass ratio of 0.78 ± 0.07 is similar to that of 2MASS J1404−3159AB. (Note that SDSS J1534+1615AB is a J-band flux-reversal binary in our input sample, but it does not have a secure orbit determination.) However, in that case there were discrepancies between models and observations due to the fact that Lyon models, either Cond here or Dusty in Dupuy et al. (2015b), predicted a much steeper drop in luminosity given the change in mass. The lack of such discrepancies for 2MASS J1404−3159AB is because despite having a similar mass ratio, its luminosity ratio is much further from unity, ${\rm{\Delta }}\mathrm{log}({L}_{\mathrm{bol}})=0.35\pm 0.09$ dex compared to 0.13 ± 0.08 dex for SDSS J1052+4422AB. The SM08 model-derived temperature for 2MASS J1404−3159A (${1400}_{-50}^{+40}$ K) is comparable to that of SDSS J1052+4422A (${1366}_{-29}^{+25}$ K), while 2MASS J1404−3159B (1190 ± 50 K) is slightly cooler than SDSS J1052+4422B (1270 ± 40 K).

      2MASS J1404−3159A is consistent with being massive enough that it would be depleted in lithium, $\mathrm{log}(\mathrm{Li}/{\mathrm{Li}}_{\mathrm{init}}\,=-{1.0}_{-0.5}^{+1.0}$ dex, although within the 1σ range includes the possibility that it has not depleted any lithium. In contrast, the Cond models predict that 2MASS J1404−3159B retains almost all of its initial lithium, $\mathrm{log}(\mathrm{Li}/{\mathrm{Li}}_{\mathrm{init}}\,=-{0.018}_{-0.006}^{+0.003}$ dex. However, the T5 spectral type of 2MASS J1404−3159B suggests that no Li i will be observable in its spectrum because monatomic lithium will be chemically depleted into LiCl and/or LiOH. No optical spectroscopy is yet available to test these model predictions of lithium depletion, or lack thereof, for this system.

      B.15. HD 130948BC (L4+L4)

      The properties of HD 130948BC and the corresponding astrophysical interpretation are discussed in detail in Dupuy et al. (2009b) and updated in Dupuy et al. (2014). Our new homogeneous analysis here results in slightly revised properties for both mass and luminosity, but the key results are unchanged. (HD 130948A does not appear in the Gaia-DR1 Tycho-Gaia Astrometric Solution catalog of parallaxes.)

      B.16. Gl 569Bab (M8.5+M9)

      In our previous work, we used Keck and HST data to measure the orbit of Gl 569Bab, finding a semimajor axis of ${95.6}_{-1.0}^{+1.1}$ mas (Dupuy et al. 2010) from which we computed a system mass of ${147}_{-8}^{+9}$ ${M}_{\mathrm{Jup}}$. We discussed why this may be discrepant with other literature measurements, the smallest of which was 90.4 ± 0.7 mas from Simon et al. (2006). Our new measurement of 93.64 ± 0.14 mas, based on a much more extensive data set with better orbital phase coverage, is still inconsistent (4.1σ) with the Simon et al. (2006) value, and it is also 2.0σ smaller than our previously published value. This results in a somewhat smaller system mass of 138 ± 7 ${M}_{\mathrm{Jup}}$, which is only 0.8σ different from our previously published value, because the error is dominated by the Hipparcos parallax (${\sigma }_{\pi }/\pi =0.016$), not our semimajor axis measurement (σa/a = 0.0016). (Gl 569A does not appear in the Gaia-DR1 Tycho-Gaia Astrometric Solution catalog of parallaxes.)

      The BHAC model-derived age of Gl 569Bab is ${440}_{-60}^{+50}$ Myr, in good agreement with our past work (${460}_{-110}^{+70}$ Myr from the Dusty models; Dupuy et al. 2010). The model-derived mass ratio is q = 0.85 ± 0.03, or 1/q = 1.17 ± 0.05. This is in good agreement with the measured mass ratio of 1/q = 1.4 ± 0.3 based on a homogeneous analysis of radial velocities by Konopacky et al. (2010). The model-derived individual masses are 75 ± 4 ${M}_{\mathrm{Jup}}$ and 64 ± 4 ${M}_{\mathrm{Jup}}$. Using our total mass and the mass ratio from Konopacky et al. (2010) results in individual masses of ${80}_{-8}^{+9}$ ${M}_{\mathrm{Jup}}$ and ${58}_{-9}^{+7}$ ${M}_{\mathrm{Jup}}$. In both cases, given our empirical substellar boundary of ≈70 ${M}_{\mathrm{Jup}}$, Gl 569Bb is likely a brown dwarf and Gl 569Ba could be a brown dwarf or a star.

      B.17. 2MASS J1534−2952AB (T4.5+T5)

      In our previous work, Liu et al. (2008) used a parallax of 73.6 ± 1.2 mas from Tinney et al. (2003) and a different orbit analysis method to determine a total dynamical mass of 59 ± 3 ${M}_{\mathrm{Jup}}$. Our CFHT parallax (first published in Dupuy & Liu 2012) is 63.0 ± 1.1 mas, putting this binary significantly farther away and thus making its semimajor axis and mass measurements larger. Liu et al. (2008) quoted a total mass of ${58.3}_{-1.8}^{+2.0}$ M at a fixed parallax of 73.6 mas. We can therefore readily compute how the total mass from that orbit analysis would have changed if we instead use our parallax, finding ${92.9}_{-2.8}^{+3.2}$ ${M}_{\mathrm{Jup}}$ at a fixed parallax of 63.0 mas. The dynamical mass we derive here with more data and a different analysis method is ${99.5}_{-0.6}^{+0.8}$ ${M}_{\mathrm{Jup}}$ at a fixed parallax of 63.0 mas, which is somewhat higher but consistent within 2σ. As discussed in Section 4.2, this is largely due to a difference in the eccentricity prior for this nearly circular orbit. After incorporating the parallax uncertainty, our newly determined total mass is 99 ± 5 ${M}_{\mathrm{Jup}}$.

      In our total-mass analysis, the SM08 models give an age of ${3.0}_{-0.5}^{+0.4}$ Gyr, which is somewhat older than the Cond age of ${2.25}_{-0.32}^{+0.29}$ Gyr. Models give consistent mass ratios of 0.96 ± 0.06 (SM08) and 0.95 ± 0.07 (Cond), which agree well with our measured mass ratio of ${0.95}_{-0.16}^{+0.13}$. Thus, properties derived from our total-mass analysis agree well with our individual-mass analysis, and the two components are coeval according to models. With individual masses of 51 ± 5 ${M}_{\mathrm{Jup}}$ and 48 ± 5 ${M}_{\mathrm{Jup}}$, both components are predicted by the Cond models to have retained ≥95% (1σ) of their initial lithium. No Li i absorption is observed in the integrated-light spectrum of 2MASS J1534−2952AB (Burgasser et al. 2003a), implying that monatomic lithium is likely chemically depleted into LiCl and/or LiOH as expected. The component temperatures derived from the Cond models are 1170 ± 50 K and ${1110}_{-50}^{+40}$ K. This provides an empirical lower limit on the effective temperature at which lithium is chemically depleted for field brown dwarfs. Finally, we note that despite being the latest-type binary in our sample, the components of 2MASS J1534−2952AB are in fact similar in mass to some of the L and early-T dwarfs in our sample (e.g., SDSS J0423−0414A, SDSS J1052+4422A, Gl 417B, and Gl 417C).

      B.18. 2MASS J1728+3948AB (L5+L7)

      Combining our total mass (${140}_{-8}^{+7}$ ${M}_{\mathrm{Jup}}$) and mass ratio (${0.93}_{-0.13}^{+0.11}$) gives individual masses of 73 ± 7 ${M}_{\mathrm{Jup}}$ for 2MASS J1728+3948A (L5 ± 1) and 67 ± 5 ${M}_{\mathrm{Jup}}$ for 2MASS J1728+3948B (L7 ± 1). In our total-mass analysis, models give mass ratios in good agreement with our measurement, and correspondingly the ages derived in our individual-mass analysis are consistent between the primary and secondary. The total-mass analysis gives the tightest age constraints of ${3.4}_{-2.1}^{+2.8}$ Gyr (SM08) and ${2.9}_{-1.1}^{+0.7}$ Gyr (Cond). 2MASS J1728+3948A is predicted to be strongly depleted of lithium, with $\mathrm{log}(\mathrm{Li}/{\mathrm{Li}}_{\mathrm{init}})=-{2.9}_{-0.7}^{+0.4}$ dex, while within 1σ, 2MASS J1728+3948B ranges from somewhat to severely depleted in lithium, $\mathrm{log}(\mathrm{Li}/{\mathrm{Li}}_{\mathrm{init}})=-{1.5}_{-0.6}^{+0.7}$ dex. Given these predictions, it would be possible for the integrated-light spectrum to show weak lithium absorption due to 2MASS J1728+3948B. The lack of a strong Li i detection (EW < 4 Å) by Kirkpatrick et al. (2000) does not rule out such a weak absorption, although it is also consistent with both components being fully depleted of lithium. In that case, this nondetection of lithium would be an indication that the secondary mass is on the higher side of our quoted 1σ intervals.

      B.19. LSPM J1735+2634AB (M7.5+L0)

      Combining our total mass (187 ± 7 ${M}_{\mathrm{Jup}}$) and mass ratio (${0.868}_{-0.025}^{+0.023}$) gives individual masses of 100 ± 4 ${M}_{\mathrm{Jup}}$ for LSPM J1735+2634A (M7.5) and 87 ± 3 ${M}_{\mathrm{Jup}}$ for LSPM J1735+2634B (L0). Having one of the more precise mass ratios, and one that is further from unity than most, LSPM J1735+2634AB provides a good test of the model-predicted mass–luminosity relation. In our total-mass analysis, BHAC models predict a mass ratio of ${0.913}_{-0.017}^{+0.019}$, which is only marginally (1.6σ) higher than what we measure. Because the components of LSPM J1735+2634AB are both likely main-sequence stars, the age is essentially unconstrained by models, which only give a 3σ lower limit of >0.6 Gyr from both total-mass and individual-mass analyses.

      B.20. 2MASS J2132+1341AB (L4.5+L8.5)

      Combining our total mass (${128}_{-8}^{+7}$ ${M}_{\mathrm{Jup}}$) and mass ratio (${0.88}_{-0.05}^{+0.04}$) gives individual masses of 68 ± 4 ${M}_{\mathrm{Jup}}$ for 2MASS J2132+1341A (L4.5 ± 1.5) and 60 ± 4 ${M}_{\mathrm{Jup}}$ for 2MASS J2132+1341B (L8.5 ± 1.5). In our total-mass analysis, models give mass ratios in good agreement with our measurement, and correspondingly the ages derived in our individual mass analysis are consistent between the primary and secondary. The total-mass analysis gives the tightest age constraints of ${1.44}_{-0.37}^{+0.26}$ Gyr (SM08) and ${1.71}_{-0.38}^{+0.28}$ Gyr (Cond). 2MASS J2132+1341A is predicted to be severely depleted of lithium, while 2MASS J2132+1341B straddles the lithium-fusion boundary within the errors but is predicted to mostly be above it and thus not strongly be depleted in lithium, $\mathrm{log}(\mathrm{Li}/{\mathrm{Li}}_{\mathrm{init}})=-{1.1}_{-0.4}^{+1.1}$. If Li i were present, then Cruz et al. (2007) would likely have detected it in their integrated-light spectrum of 2MASS J2132+1341, although they do not quote EW limits. Their lack of a detection would be more consistent with the model predictions if 2MASS J2132+1341B is on the more massive end of our measurement uncertainties.

      B.21. 2MASS J2140+1625AB (M8+L0.5)

      Our measured mass ratio for 2MASS J2140+1625AB is quite low (${0.60}_{-0.08}^{+0.07}$), such that it divides up our total mass (${183}_{-17}^{+14}$ ${M}_{\mathrm{Jup}}$) into a quite massive M8 primary (${114}_{-12}^{+10}$ ${M}_{\mathrm{Jup}}$) and a rather low-mass L0.5 secondary (${69}_{-9}^{+8}$ ${M}_{\mathrm{Jup}}$). This is strongly disfavored by models, mainly because the primary is much less luminous than expected for such a high mass. Our total-mass analysis using BHAC models predicts a mass ratio of ${0.882}_{-0.024}^{+0.028}$, 3.8σ discrepant with our measurement. Interestingly, both our total-mass and individual-mass analyses end up with an essentially identical total mass as we measure. The individual-mass analysis demonstrates how this can be the case despite such a discrepant model-derived mass ratio: the primary mass after performing rejection sampling was 1.7σ lower than what we input, and the secondary mass was 2.0σ higher than what we input.

      This apparent discrepancy could be jointly due to two smaller (≈1.5σ) discrepancies in the measured mass ratio and parallax. 2MASS J2140+1625 has one of the less precise mass ratios in our sample and a 3% parallax uncertainty that leads to an atypically large total mass error (${}_{-9}^{+8} \% ).$ Therefore, higher-precision data would be useful in determining if 2MASS J2140+1625AB is truly an astrophysical outlier.

      B.22. 2MASS J2206−2047AB (M8+M8)

      In our previous work, we used a parallax of 37.5 ± 3.4 mas from Costa et al. (2006) to determine a dynamical total mass and model-derived properties (Dupuy et al. 2009a). Our CFHT parallax of 35.7 ± 1.2 mas was first presented in Dupuy & Liu (2012), and the updated analysis here gives 35.8 ± 1.0 mas, which allows for a more precise total mass than in Dupuy et al. (2009a), as well as individual masses from our photocenter analysis. Our total mass is ${188}_{-17}^{+16}$ ${M}_{\mathrm{Jup}}$, consistent with our previous measurement of ${160}_{-30}^{+50}$ ${M}_{\mathrm{Jup}}$ (Dupuy et al. 2009a). Our measured mass ratio of ${0.84}_{-0.10}^{+0.09}$, consistent with the BHAC model-derived value of ${0.986}_{-0.029}^{+0.028}$, gives component masses of ${102}_{-11}^{+10}$ ${M}_{\mathrm{Jup}}$ and ${86}_{-10}^{+8}$ ${M}_{\mathrm{Jup}}$. In both total-mass and individual-mass analyses, rejection sampling results in slightly higher secondary masses and lower primary masses in the output distributions (96 ${M}_{\mathrm{Jup}}$ and 95 ${M}_{\mathrm{Jup}}$), but they are consistent with the inputs to within ≤1σ. The masses and luminosities of 2MASS J2206−2047AB are consistent with this being a pair of main-sequence stars, with BHAC models giving a 3σ lower limit on the age of >0.27 Gyr.

      B.23. DENIS J2252−1730AB (L4+T3.5)

      Combining our total mass (101 ± 7 ${M}_{\mathrm{Jup}}$) and mass ratio (${0.70}_{-0.09}^{+0.08}$) gives individual masses of 59 ± 5 ${M}_{\mathrm{Jup}}$ for DENIS J2252−1730A (L4.0 ± 1.0) and 41 ± 4 ${M}_{\mathrm{Jup}}$ for DENIS J2252−1730B (T3.5 ± 0.5). Therefore, both components are unambiguously substellar. In our total-mass analysis, the SM08 models prefer a somewhat lower mass ratio (${0.57}_{-0.05}^{+0.03}$), consistent with our measurement within the uncertainties, while the Cond models predict a somewhat higher mass ratio (${0.71}_{-0.04}^{+0.05}$). Both sets of models give consistent ages, with the total mass analysis providing the tightest age constraints of ${1.11}_{-0.22}^{+0.19}$ Gyr (SM08) and ${1.10}_{-0.18}^{+0.15}$ Gyr (Cond). According to the Cond models, DENIS J2252−1730B is predicted to have retained nearly all of its lithium, while DENIS J2252−1730A straddles the lithium fusion boundary within errors, $\mathrm{log}(\mathrm{Li}/{\mathrm{Li}}_{\mathrm{init}})\,=-{2.1}_{-1.5}^{+0.7}$ dex. DENIS J2252−1730B is quite cool (${1230}_{-60}^{+50}$ K, Cond) and thus may not be expected to display Li i in absorption due to being chemically depleted. DENIS J2252−1730A, on the other hand, is an L2 dwarf that should be amenable to lithium-absorption measurement. Reid et al. (2008b) obtained an integrated-light optical spectrum for this relatively bright binary and did not report lithium absorption for this source. This suggests that DENIS J2252−1730A is indeed above the lithium-fusion limit and thus its true mass may be in the higher range of our quoted 1σ mass interval.

      Footnotes

      • Data presented herein were obtained at the W.M. Keck Observatory, which is operated as a scientific partnership among the California Institute of Technology, the University of California, and the National Aeronautics and Space Administration. The Observatory was made possible by the generous financial support of the W.M. Keck Foundation.

      • † 

        Based on data obtained with WIRCam, a joint project of CFHT, Taiwan, Korea, Canada, France, at the Canada–France–Hawaii Telescope, which is operated by the National Research Council of Canada, the Institute National des Sciences de l'Univers of the Centre National de la Recherche Scientifique of France, and the University of Hawaii.

      • ‡ 

        Based on observations made with the NASA/ESA Hubble Space Telescope, obtained at the Space Telescope Science Institute, which is operated by the Association of Universities for Research in Astronomy, Inc., under NASA contract NAS 5-26555. These observations are associated with programs GO-11593, GO-12317, and GO-12661.

      • We note the following significant updates to Table 1 of Burgasser et al. (2007): 2MASS J0652307+471034AB is listed as a 2.0 au binary (I. N. Reid et al. 2008, in preparation), but Reid et al. (2006b), Konopacky et al. (2010), and Keck LGS AO observations of our own confirm that it is single; SDSS J233558.51–001304.1AB is much more distant (160 ± 30 pc) and thus wider (9.0 ± 1.7 au) than previously estimated given that its spectral type is M7 (West et al. 2008) and not early- to mid-L.

      • Our spectral-type cuts excluded the late-type binaries 2MASSW J1225543–273947AB (T6) and 2MASS J15530228+1532369AB (T7). At earlier types, we excluded LSPM J1314+1320AB based on an estimated spectral type of M6 from Law et al. (2006), but note that it has since been updated to M7 (Lépine et al. 2009; Dupuy et al. 2016). Other notable systems near the boundary of our spectral type cutoff and thus not included here are L726-8AB a.k.a. Gl 69AB (M5.5+M6; Geyer et al. 1988; Kirkpatrick et al. 1991), L789-6ABC a.k.a. Gl 866ABC (M5+M5.5:+M6.5:; Henry et al. 1994; Delfosse et al. 2000), and GJ 1245ABC (M5.5+M5.5+M6.5:; Kirkpatrick et al. 2012; Benedict et al. 2016).

      • By focusing on nearby systems, the only binaries with projected separations ≤6 au that we excluded were 2MASS J16000548+1708328AB (61 pc; Bouy et al. 2003) and binaries in Taurus and Upper Scorpius (Kraus et al. 2005, 2006). The most distant binary in our sample is LP 415-20AB at 39.7 ± 1.1 pc, and its original estimated distance was actually 30 ± 5 pc (Siegler et al. 2003).

      • Programs GO-11593, GO-12317, and GO-12661 (PI Dupuy).

      • 10 
      • 11 

        Here, in the absence of radial velocity information, there is a 180° ambiguity in Ω, and consequently ω and λref, such that either Ω or ${\rm{\Omega }}+180^\circ $ is actually the ascending node while the other is the descending node. Therefore, we follow the standard convention that 0° ≤ Ω ≤ 180°. If future radial velocities show that ${\rm{\Omega }}+180^\circ $ is actually the ascending node then 180° should be added to our reported values of Ω, ω, and λref.

      • 12 
      • 13 

        To be clear about the terminology we are using, "star formation rate" refers to the number of stars and brown dwarfs that were formed in the Galactic disk as a function of time. "Age distribution" refers to the local, present day number of stars and brown dwarfs as a function of age.

      • 14 
      • 15 
      • 16 

        We quote the mass implied by their given period and semimajor axis, because their own quoted mass seems to assume a conversion factor of 1000 ${M}_{\mathrm{Jup}}$/M, not the correct 1048 ${M}_{\mathrm{Jup}}$/M.

      Please wait… references are loading.
      10.3847/1538-4365/aa5e4c